0% found this document useful (0 votes)
10 views

optical-microring-resonators-theory-techniques-and-applications_compress

The document is a comprehensive resource on Optical Microring Resonators, detailing their theory, techniques, and applications. It includes various chapters covering elements, analytical models, coupled optical filters, nonlinear optics applications, and active photonic applications. The book is part of a series in optics and optoelectronics and is authored by Vien Van from the University of Alberta.

Uploaded by

Leonrdo Frezz
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views

optical-microring-resonators-theory-techniques-and-applications_compress

The document is a comprehensive resource on Optical Microring Resonators, detailing their theory, techniques, and applications. It includes various chapters covering elements, analytical models, coupled optical filters, nonlinear optics applications, and active photonic applications. The book is part of a series in optics and optoelectronics and is authored by Vien Van from the University of Alberta.

Uploaded by

Leonrdo Frezz
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 288

Optical Microring

Resonators
Theory, Techniques,
and Applications
SERIES IN OPTICS AND OPTOELECTRONICS
Series Editors: E Roy Pike, Kings College, London, UK
Robert G W Brown, University of California, Irvine, USA

Recent titles in the series


Optical Microring Resonators: Theory, Techniques, and Applications
V. Van
Optical Compressive Imaging
Adrian Stern
Singular Optics
Gregory J. Gbur
The Limits of Resolution
Geoffrey de Villiers and E. Roy Pike
Polarized Light and the Mueller Matrix Approach
José J Gil and Razvigor Ossikovski
Light—The Physics of the Photon
Ole Keller
Advanced Biophotonics: Tissue Optical Sectioning
Ruikang K Wang and Valery V Tuchin (Eds.)
Handbook of Silicon Photonics
Laurent Vivien and Lorenzo Pavesi (Eds.)
Microlenses: Properties, Fabrication and Liquid Lenses
Hongrui Jiang and Xuefeng Zeng
Laser-Based Measurements for Time and Frequency Domain
Applications: A Handbook
Pasquale Maddaloni, Marco Bellini, and Paolo De Natale
Handbook of 3D Machine Vision: Optical Metrology and Imaging
Song Zhang (Ed.)
Handbook of Optical Dimensional Metrology
Kevin Harding (Ed.)
Biomimetics in Photonics
Olaf Karthaus (Ed.)
Optical Properties of Photonic Structures: Interplay of Order
and Disorder
Mikhail F Limonov and Richard De La Rue (Eds.)
Nitride Phosphors and Solid-State Lighting
Rong-Jun Xie, Yuan Qiang Li, Naoto Hirosaki, and Hajime Yamamoto
Optical Microring
Resonators
Theory, Techniques,
and Applications

V. Van
University of Alberta
Edmonton, Canada
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2017 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works

Printed on acid-free paper


Version Date: 20161004

International Standard Book Number-13: 978-1-4665-5124-4 (Hardback)

This book contains information obtained from authentic and highly regarded sources. Reasonable
efforts have been made to publish reliable data and information, but the author and publisher cannot
assume responsibility for the validity of all materials or the consequences of their use. The authors and
publishers have attempted to trace the copyright holders of all material reproduced in this publication
and apologize to copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we may rectify in any
future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information stor-
age or retrieval system, without written permission from the publishers.

For permission to photocopy or use material electronically from this work, please access www.copy-
right.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222
Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that pro-
vides licenses and registration for a variety of users. For organizations that have been granted a pho-
tocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation without intent to infringe.

Library of Congress Cataloging‑in‑Publication Data

Names: Van, Vien, 1971- author.


Title: Optical microring resonators : theory, techniques, and applications /
Vien Van.
Other titles: Series in optics and optoelectronics (CRC Press)
Description: Boca Raton, FL : CRC Press, Taylor & Francis Group, [2017] |
Series: Series in optics and optoelectronics
Identifiers: LCCN 2016040265| ISBN 9781466551244 (hardback ; alk. paper) |
ISBN 1466551240 (hardback ; alk. paper)
Subjects: LCSH: Microresonators (Optoelectronics) | Optoelectronic devices. |
Optical wave guides. | Nonlinear optics.
Classification: LCC TK8360.M53 V36 2017 | DDC 621.381/045--dc23
LC record available at https://lccn.loc.gov/2016040265

Visit the Taylor & Francis Web site at


http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Contents

Preface ix

1 Elements of an Optical Microring Resonator. . . . . . . . . . . . . . . . 1


1.1 Dielectric Optical Waveguides. . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 The vectorial wave equations. . . . . . . . . . . . . . . . . 4
1.1.2 The EIM and solutions of the
one-dimensional slab waveguide. . . . . . . . . . . . . . 10
1.1.3 Waveguide dispersion. . . . . . . . . . . . . . . . . . . . . . . 13
1.1.4 Propagation loss. . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.2 Optical Modes in Bent Dielectric Waveguides . . . . . . . . . 18
1.2.1 Conformal transformation of bent waveguides. . .18
1.2.2 Resonant modes in microdisks and
microrings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.2.3 Full-vectorial analysis of bent waveguides. . . . . 30
1.3 Coupling of Waveguide Modes in Space . . . . . . . . . . . . . 35
1.3.1 The coupled mode equations. . . . . . . . . . . . . . . . 35
1.3.2 Solution of the coupled mode equations. . . . . . . 39
1.4 Fabrication of Microring Resonators. . . . . . . . . . . . . . . . . 44
1.5 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2 Analytical Models of a Microring Resonator. . . . . . . . . . . . . . . 53


2.1 Resonance Spectrum of an Uncoupled Microring
Resonator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.2 Power Coupling Description of a Microring Resonator. 57
2.2.1 The add-drop microring resonator . . . . . . . . . . . 58
2.2.2 The all-pass microring resonator. . . . . . . . . . . . . . 67
2.2.3 Phase and GD responses of a microring
resonator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
2.3 Energy Coupling Description of a Microring Resonator. . . . 75
2.4 Relationship between Energy Coupling and Power
Coupling Formalisms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
2.5 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

v
vi Optical Microring Resonators

3 Coupled Microring Optical Filters . . . . . . . . . . . . . . . . . . . . . . . 87


3.1 Periodic Arrays of Microring Resonators. . . . . . . . . . . . . 88
3.1.1 Periodic arrays of APMRs. . . . . . . . . . . . . . . . . . . 89
3.1.2 Periodic arrays of ADMRs. . . . . . . . . . . . . . . . . . . 93
3.1.3 Arrays of serially coupled microring
resonators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.2 Transfer Functions of Coupled Microring
Optical Filters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
3.3 Cascaded All-Pass Microring Filters. . . . . . . . . . . . . . . . . 103
3.4 Serially Coupled Microring Filters . . . . . . . . . . . . . . . . . . 107
3.4.1 Energy coupling analysis of serially
coupled microring filters. . . . . . . . . . . . . . . . . . . . 107
3.4.2 Energy coupling synthesis of serially
coupled microring filters. . . . . . . . . . . . . . . . . . . . 112
3.4.3 Power coupling analysis of serially
coupled microring filters. . . . . . . . . . . . . . . . . . . . 116
3.4.4 Power coupling synthesis of serially
coupled microring filters. . . . . . . . . . . . . . . . . . . . 119
3.5 Parallel Cascaded ADMRs. . . . . . . . . . . . . . . . . . . . . . . . 122
3.6 Parallel Cascaded Microring Doublets. . . . . . . . . . . . . . . 126
3.7 2D Networks of CMRs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
3.7.1 Energy coupling analysis of 2D CMR
networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
3.7.2 Energy coupling synthesis of 2D CMR
networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
3.7.3 Power coupling analysis of 2D CMR
networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
3.8 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

4 Nonlinear Optics Applications of Microring Resonators . . . . 163


4.1 Nonlinearity in Optical Waveguides . . . . . . . . . . . . . . . . 163
4.1.1 Intensity-dependent nonlinearity . . . . . . . . . . . . 163
4.1.2 Free carrier-induced nonlinear effects . . . . . . . . 166
4.1.3 Wave propagation in a nonlinear optical
waveguide. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.1.4 Free carrier-induced nonlinear effects in an
optical waveguide . . . . . . . . . . . . . . . . . . . . . . . . . 176
4.1.5 FWM in a nonlinear optical waveguide. . . . . . . 183
4.2 Optical Bistability and Instability in a
Nonlinear Microring Resonator. . . . . . . . . . . . . . . . . . . . . 191
Contents vii

4.2.1SPM and enhanced nonlinearity in a


microring resonator. . . . . . . . . . . . . . . . . . . . . . . . 192
4.2.2 Bistability and self-pulsation in a microring
resonator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
4.2.3 Free carrier effects and self-pulsation in a
microring resonator. . . . . . . . . . . . . . . . . . . . . . . 200
4.3 All-Optical Switching in a Nonlinear Microring
Resonator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
4.3.1 Enhanced nonlinear switching in a
microring resonator. . . . . . . . . . . . . . . . . . . . . . . . 213
4.3.2 Self-switching of a pulse in a nonlinear
microring resonator. . . . . . . . . . . . . . . . . . . . . . . . 217
4.3.3 Pump-and-probe switching in a nonlinear
microring resonator. . . . . . . . . . . . . . . . . . . . . . . 223
4.4 FWM in a Nonlinear Microring Resonator . . . . . . . . . . 226
4.5 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237

5 Active Photonic Applications of Microring Resonators. . . . . . 241


5.1 Mechanisms for Tuning Microring Resonators. . . . . . . . 241
5.1.1 The thermo-optic effect. . . . . . . . . . . . . . . . . . . . . 243
5.1.2 The electro-optic effect . . . . . . . . . . . . . . . . . . . . . 247
5.1.3 Free carrier dispersion. . . . . . . . . . . . . . . . . . . . . 250
5.2 Dynamic Response of a Microring Modulator . . . . . . . 253
5.2.1 Dynamic energy-coupling model of a
microring modulator. . . . . . . . . . . . . . . . . . . . . . 254
5.2.2 Small-signal analysis. . . . . . . . . . . . . . . . . . . . . . 255
5.2.2.1 Transfer function of an APMR
modulator. . . . . . . . . . . . . . . . . . . . . . . 258
5.2.2.2 Modulation efficiency. . . . . . . . . . . . . 258
5.2.2.3 Electrical bandwidth. . . . . . . . . . . . . . 260
5.2.2.4 Small-signal step response. . . . . . . . . 263
5.3 Large-Signal Response of a Microring Modulator. . . . 264
5.3.1 Large-signal step response. . . . . . . . . . . . . . . . . 264
5.3.2 Large-signal response under sinusoidal
modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
5.3.3 Intermodulation products. . . . . . . . . . . . . . . . . . . 274
5.4 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275

Index 279
Preface

Since their early development in the 1990s, optical microring


­resonators have become one of the most important elements in
integrated optics technology. These simple but versatile structures
have found myriad applications in filters, sensors, lasers, nonlinear
optics, ­optomechanics, and more recently, quantum optics. Indeed,
it is difficult to imagine an integrated photonic system today that
does not utilize a microring resonator as part of the circuit architec-
ture. Given the diversity and ubiquitousness of their applications,
the need arises for a systematic review of the technology that can
serve as a reference source for engineers and researchers working
in the broader field of integrated photonics.
The objective of this book is to provide a concise treatment
of the theory, principles and techniques of microring resonator
devices, and their applications. It is intended for graduate students
and researchers who wish to familiarize themselves with the tech-
nology and acquire sufficient knowledge to enable them to design
microring devices for their own applications of interest. Rapid
advances in integrated optics and microfabrication technologies
have enabled microring devices with increasingly more sophisti-
cated designs and superior performance to be developed. However,
the underlying working principles of these devices remain the same
for the most part. Thus, while attempts are made to highlight some
of the recent advances in microring technology, the main focus of
the book is to provide a detailed treatment of the underlying theory
and techniques for modeling and designing microring devices so
that the reader can readily apply the knowledge to their own appli-
cations. Toward this aim, numerous numerical examples are given
to help illustrate the application of these techniques, as well as to
demonstrate what can theoretically be achieved with these devices.
Many of the examples are based on the silicon-on-insulator mate-
rial system, a choice motivated by the growing prevalence of silicon
photonics in integrated optics technology.
The book is divided into five chapters. While a basic familiarity
with optics is assumed, a brief review of the concepts essential to

ix
x Optical Microring Resonators

the understanding of microring resonators and their applications


will be given in the relevant chapters. In particular, Chapter 1 will
give a review of the theory of optical waveguides, whispering gal-
lery modes, and coupled waveguide systems, which are the basic
­elements constituting any microring device. Chapter 2 develops the
basic formalisms used to analyze simple microring resonators cou-
pled to one or more waveguides. Chapter 3 is devoted to the analysis
and design of coupled microring resonators for filter applications.
Nonlinear optics and active photonic applications of microring
­resonators are the subjects of Chapters 4 and 5, respectively.
This book is born out of my research on microring resonators
which began at the University of Waterloo, Ontario, Canada, in the
late 1990s and subsequently at the University of Maryland, College
Park, in the early 2000s. I have in particular benefited tremendously
from the mentorship of Professor Ping-Tong Ho as well as from my
­colleagues at the Laboratory for Physical Sciences at the University
of Maryland. The contributions of my graduate students at the
University of Alberta have also helped shape a large part of this
book, with special acknowledgment to Ashok Prabhu Masilamani,
Alan Tsay, Daniel Bachman, Guangcan Mi, and Siamak Abdollahi.
I also acknowledge the assistance of many students in the prepara-
tion and editing of the book, with special thanks to Jocelyn Bachman,
Guangcan Mi, Yang Ren, and Daniel Bachman for reviewing and
proofreading parts of the manuscript.

V. Van
University of Alberta
September 2016
Chapter 1

Elements of an Optical
Microring Resonator

An optical microring resonator is an integrated optic traveling wave


resonator constructed by bending an optical waveguide to form a
closed loop, typically of a circular or racetrack shape. Light propa-
gating in the microring waveguide interferes with itself after every
trip around the ring. When the roundtrip length is exactly equal to
an integer multiple of the guided wavelength, constructive interfer-
ence of light occurs which gives rise to sharp resonances and large
intensity buildup inside the microring. The high wavelength selec-
tivity, strong dispersion, large field enhancement, and high quality
factor are important characteristics which make microring resona-
tors extremely versatile and useful for a wide range of applications
in optical communication, signal processing, sensing, nonlinear
optics and, more recently, quantum optics.
Microring and microdisk resonators were first proposed by
Marcatili in 1969 for realizing channel dropping filters based on
planar optical waveguides (Marcatili 1969b). Similar traveling wave
filters based on microwave striplines had been proposed and stud-
ied earlier by Coale (1956). However, it was not until the late 1990s
that advances in the microfabrication technology for integrated
photonic devices enabled optical microring resonators to be real-
ized with high quality factors. Since then, microring and microdisk
resonators have been demonstrated for a wide range of applications
in various material systems such as silicon-on-insulator (SOI), III–V
semiconductors, glass, and polymers.
Broadly defined, an optical resonator is a structure which con-
fines light in all spatial directions. In a microring resonator, this is
achieved in two ways: transversely by the refractive index contrast
of the dielectric waveguide used to form the microring and longi-
tudinally by the periodic boundary condition imposed by the ring
or racetrack structure. The propagation characteristics of the dielec-
tric waveguide, and especially those of the curved waveguide,
thus have a major impact on the characteristics of the microring

1
2 Optical Microring Resonators
(a) (b)
Bus waveguide Microring Bus waveguide Microdisk

Figure 1.1 Schematic of (a) a microring resonator coupled to two bus


(or access) waveguides and (b) a microdisk resonator coupled to one bus
waveguide.

resonator. In addition, the manner by which light is coupled into


and out of the resonator is also of practical importance. Typically
this is achieved via evanescent field coupling between the micror-
ing waveguide and one or two external straight waveguides, called
access or bus waveguides, by which light is coupled into or out of
the resonator. Figure 1.1 illustrates a microring resonator coupled to
two bus waveguides and its microdisk variant coupled to a single
waveguide.
This chapter provides a brief review of the basic elements con-
stituting a microring resonator, with a view on how their designs
and properties may affect the performance of the resonator. Section
1.1 reviews the theory of planar dielectric waveguides and the wave
equations governing optical mode propagation. Section 1.2 looks at
light propagation in curved waveguides and examines the proper-
ties of whispering gallery modes in microdisk and microring reso-
nators. Section 1.3 develops a formalism for analyzing the coupling
of waveguide modes in space, which is useful for designing eva-
nescent wave couplers for coupling light into and out of a micror-
ing resonator. Finally, Section 1.4 provides an overview of standard
fabrication processes for microring devices and highlights several
important issues relevant to the practical implementation of these
devices.

1.1 Dielectric Optical Waveguides


Dielectric optical waveguides are the basic structures for confin-
ing and guiding light in well-defined discrete modes in photonic
integrated circuits (PICs). A planar dielectric waveguide consists
of a core of refractive index n1 embedded in other dielectric layers
Elements of an Optical Microring Resonator 3

of lower refractive indices. Light is confined within the core due


to total internal reflection at the interfaces between the high-index
core and the lower-index cladding media. The degree of confine-
ment increases with the refractive index contrast between the core
and cladding. For a waveguide with a uniform cladding of refrac-
tive index n2, the index contrast is defined as

n12 − n22 n1 − n2
∆n = ≈ . (1.1)
2n12 n1

The approximation in the above formula is good for low-index con-


trast waveguides. Some common waveguide material systems and
their index contrasts are listed in Table 1.1. Typically, the index con-
trast ranges from about 1% for weakly confined waveguides based
on doped silica materials, to over 40% for strongly confined semi-
conductor waveguides. In general, high-index contrast (or high-Δn)
waveguides are desirable for the miniaturization of PICs since they
have smaller dimensions and provide stronger confinement of
light, which enables sharp waveguide bends to be realized with low
bending loss. On the other hand, polarization-dependent effects
and scattering loss also tend to be more pronounced in high-Δn
waveguides.
Two basic optical waveguide structures are shown in Figure 1.2:
the rib (or ridge) waveguide and the rectangular strip waveguide.
From the fabrication point of view, these two structures differ only
by the etch depth in defining the waveguide core: in a rib waveguide,
the core is etched only to a depth h leaving a residual high-index

Table 1.1 Refractive Indices of Some Common


Integrated Optic Waveguide Materials
Refractive Index Index Contrasta
Core Material at λ = 1.55 μm Δn (%)
Doped silica 1.45–1.5 0.7–4
Polymers 1.45–1.7 0.7–14
SiOxNy 1.45–2.0 0.7–24
SiNx 2.0–2.3 24–30
III–V (InP, GaAs) 3.16, 3.4 40, 41
Si 3.47 41
a Index contrast assuming SiO2 cladding with refractive
index n2 = 1.44.
4 Optical Microring Resonators
(a) (b)
Over-cladding n2 Over-cladding n2
h
n1 Core n1 t
Core layer t
y
n3 n3
Under-cladding x Under-cladding

1.2
1.0
0.6
0.8
0.6 0.4
0.4
y (µm)

0.2

y (µm)
0.2
0 0
–0.2
–0.2
–0.4
–0.6 –0.4
–1 –0.5 0 0.5 1 –0.4 –0.2 0 0.2 0.4 0.6
x (µm) x (µm)

Figure 1.2 Schematic and Ey-field distribution of the quasi-TM mode of


(a) a rib (or ridge) waveguide and (b) a rectangular strip waveguide.

layer of thickness t, whereas in a strip waveguide, the core layer is


completely etched through. However, the modal ­characteristics of
the two waveguides are quite different, as shown in Figure 1.2. In
the rib waveguide, the residual high-index layer causes the mode
to expand laterally. Due to the weak lateral ­confinement, rib wave-
guides tend to suffer from larger bending loss than strip wave-
guides. As a result, microring resonators are typically designed
using strip waveguides to minimize radiation loss due to bending.
Since the performance of a PIC (photonic integrated circuit)
depends critically on the properties of the waveguide modes, it is
important to obtain a detailed analysis of the propagation charac-
teristics of light in the optical waveguide. In the next section, we
will derive the wave equation and its various approximations for
describing electromagnetic wave propagation in a planar ­dielectric
waveguide. A review of the main methods for solving these
­equations will also be given.

1.1.1 The vectorial wave equations


The field distributions of a waveguide mode and its associated
propagation constant are determined by solving an eigenvalue
problem formulated in terms of either the transverse electric (TE)
Elements of an Optical Microring Resonator 5

field or transverse magnetic (TM) field. We consider a dielectric


waveguide oriented along the z-axis and characterized by a trans-
verse index profile n(x, y). Under the assumption that the transverse
index profile is invariant along the propagation axis (the z-axis), the
electric and magnetic field distributions of a waveguide mode can
be expressed as

E( x , y , z) = E(x , y )e − jβz , (1.2)

H( x , y , z) = H( x , y )e − jβz , (1.3)

where β is the propagation constant of the waveguide mode and the


time dependence ejωt is assumed and suppressed. For the purpose
of modal analysis, it is convenient to further decompose the fields
E and H into a transverse component and a longitudinal compo-
nent as follows:

E( x , y ) = E t ( x , y ) + Ez ( x , y )z , (1.4)

H( x , y ) = Ht ( x , y ) + H z ( x , y )z . (1.5)

The electric and magnetic fields E and H satisfy Maxwell’s


equations,

∇ × E = − jωµ 0 H, (1.6)

∇ × H = jωε 0 n2 ( x , y )E, (1.7)

where ε0 and μ0 are the electric permittivity and magnetic perme-


ability, respectively, of vacuum. By taking the curl of Equation 1.6
and using Equation 1.7 to eliminate ∇ × H from the resulting equa-
tion, we get

∇ × ∇ × E = n2 k 2 E , (1.8)

where k = ω/c. With the help of the vector identity ∇ × ∇ × E =


∇(∇ ⋅ E ) − ∇ 2E , we can write Equation 1.8 as

∇ 2E + n2 k 2E = ∇(∇ ⋅ E ). (1.9)

Substituting E( x , y , z) = E( x , y )e − jβz into the above equation and


making use of the field decomposition in Equation 1.4, we obtain
6 Optical Microring Resonators

 ∂E 
∇ 2t E t + (n2 k 2 − β 2 )E t = ∇ t (∇ t ⋅ E t ) + ∇ t  z  , (1.10)
 ∂z 

where ∇ t = xˆ (∂/∂x) + yˆ (∂/∂y ). In the absence of free charge, Gauss’s


law gives

∂Ez
∇ ⋅ ( n 2 E) = ∇ t ⋅ ( n 2 E t ) + n 2 = 0, (1.11)
∂z

from which we get

∂Ez 1
= − 2 ∇ t ⋅ (n2E t ). (1.12)
∂z n

Upon substituting the above expression into Equation 1.10, we obtain


the vectorial wave equation in terms of the transverse electric field,

1 
∇ 2t E t + (n2 k 2 − β 2 )E t = ∇ t (∇ t ⋅ E t ) − ∇ t  2 ∇ t ⋅ (n 2E t ) . (1.13)
n 

Equation 1.13 is an eigenvalue problem whose solution gives the


transverse field distribution Et of an optical mode and its propa-
gation constant β. The effective index of the waveguide mode is
defined as neff = β/k.
The terms on the right-hand side of Equation 1.13 account for
the polarization coupling between the transverse field components
Ex and Ey. Thus, in general, the mode of an optical waveguide is
hybrid or vectorial in nature, that is, it contains both Ex and Ey com-
ponents of the electric field. We can write Equation 1.13 in the form
of an eigenvalue matrix equation as (Xu et al. 1994)

 Pxx Pxy   Ex   Ex 
P    = β2   , (1.14)
 yx Pyy  Ey  Ey 

where the operators in the matrix are given by

∂  1 ∂ 2  ∂ 2 Ex
Pxx Ex = 
∂x  n ∂x
2 (
n Ex  +
 ∂y
2 )
+ n 2 k 2 Ex , (1.15)

∂ 2 Ey ∂ 1 ∂ 2 
Pyy Ey =
∂x 2
+ 
∂y  n 2 ∂y
( )
n Ey  + n 2 k 2 Ey , (1.16)

Elements of an Optical Microring Resonator 7

∂  1 ∂ 2  ∂ Ey
2
Pxy Ey =  2 (
n Ey  − ) , (1.17)
∂x  n ∂y  ∂y ∂x

∂  1 ∂ 2  ∂ 2 Ex
Pyx Ex =
∂y  n2 ∂x
( )
n Ex  −
 ∂x ∂y
. (1.18)

It is apparent from Equation 1.14 that the operators Pxy and Pyx give
rise to polarization coupling effects. For rectangular waveguides
with low to moderate index contrasts, the two lowest-order modes
are predominantly linearly polarized along either the principal
x- or y-axis. It is often a good approximation to neglect the minor
field component of each mode and consider the mode to be either
quasi-TE with major field component Ex, or quasi-TM with major
field component Ey. Under this semi-vectorial approximation, the
cross-polarization coupling terms in Equation 1.14 are neglected so
that the equations governing the major field components become

∂  1 ∂ 2  ∂ 2 Ex
Pxx Ex =
∂x  n2 ∂x
( )
n Ex  +
 ∂y
2
+ n 2 k 2 Ex = βTE
2
Ex , (quasi-TE)

(1.19)

∂ 2 Ey ∂ 1 ∂ 2 
Pyy Ey =
∂x 2
+ 
∂y  n 2 ∂y
( )
n Ey  + n 2 k 2 Ey = βTM
2
Ey . (quasi-TM)

(1.20)
For low-index contrast waveguides, one may further neglect
the spatial index variation in the square bracket terms in the above
equations. Under this approximation, the TE and TM modes become
identical and are described by the scalar wave equation

∂2E ∂2E
Pxx E = Pyy E = + + n2 k 2 E = β 2 E. (1.21)
∂x 2 ∂y 2

Figure 1.3 shows the electric field distributions of the two


­lowest-order modes in an SOI strip waveguide consisting of a Si
core of 250 nm thickness and 400 nm width embedded in a SiO2
cladding. Both the semi-vectorial and full-vectorial solutions of the
modes are shown for comparison. Also shown are the effective indi-
ces of the modes at the 1.55 μm wavelength. We see that the semi-
vectorial and full-vectorial solutions give similar field distributions
8 Optical Microring Resonators
Semi-vectorial solution Full-vectorial solution

(a) (b) Major field component Minor field component


neff = 2.411 neff = 2.345
0.6 Ex 0.6 Ex 0.6 Ey
0.4 0.4 0.4

y (µm)

y (µm)
y (µm)

0.2 0.2 0.2


0 0 0
–0.2 –0.2 –0.2
–0.4 –0.4 –0.4
–0.4 –0.2 0 0.2 0.4 0.6 –0.4 –0.2 0 0.2 0.4 0.6 –0.4 –0.2 0 0.2 0.4 0.6
x (µm) x (µm) x (µm)

(c) neff = 2.013 (d) neff = 1.914

0.6 Ey 0.6 Ey 0.6 Ex


0.4 0.4 0.4
y (µm)

y (µm)
0.2
y (µm)

0.2 0.2

0 0 0

–0.2 –0.2 –0.2

–0.4 –0.4 –0.4


–0.4 –0.2 0 0.2 0.4 0.6 –0.4 –0.2 0 0.2 0.4 0.6 –0.4 –0.2 0 0.2 0.4 0.6
x (µm) x (µm) x (µm)

Figure 1.3 Semi-vectorial and full-vectorial solutions of the two lowest-


order modes at 1.55 μm wavelength of an SOI strip waveguide (Si core of
dimensions 400 × 250 nm2 embedded in a SiO2 cladding): (a, b) TE mode,
(c, d) TM mode.

for the major field component of each mode, although there is a dif-
ference of about 3–5% between the effective index values.
We can also formulate the wave equation in terms of the trans-
verse magnetic field Ht. One advantage of solving for the optical
mode in terms of the magnetic field is that the fields Hx and Hy are
continuous across all dielectric boundaries. By taking the curl of
Equation 1.7 and using Equation 1.6 to eliminate ∇ × E , we get

1
∇ 2H + n2 k 2H = − ∇n2 × (∇ × H ). (1.22)
n2

Substituting Equations 1.3 and 1.5 into the above equation, we


obtain the following vectorial wave equation in terms of the trans-
verse magnetic field,

1
∇ 2t Ht + (n2 k 2 − β 2 )Ht = − 2
(∇ t n2 ) × (∇ t × Ht ). (1.23)
n
Elements of an Optical Microring Resonator 9

The above equation can be written explicitly in a component form


as (Xu et al. 1994)

Qxx Qxy   H x   Hx 
Q    = β2   , (1.24)
 yx Qyy   H y   Hy 

where

∂ 2 Hx ∂  1 ∂H x 
Qxx H x = 2
+ n2   + n2 k 2 Hx , (1.25)
∂x ∂y  n ∂y 
2

2
∂  1 ∂H y  ∂ H y
Qyy H y = n2 + + n2 k 2 H y , (1.26)
∂x  n2 ∂x  ∂y 2

∂2 Hy ∂  1 ∂H y 
Qxy H y = − n2 , (1.27)
∂x ∂y ∂y  n2 ∂x 

∂  1 ∂H x  ∂ 2 H x
Qyx H x = n2 + . (1.28)
∂x  n2 ∂y  ∂y ∂x

In general, the vectorial wave equations (1.13) and (1.23) do not


have analytical solutions and must be solved numerically. Many
efficient numerical techniques have been developed for solving
these equations for waveguides with arbitrary cross-sections and
index profiles, the most popular ones being the finite difference
method (Xu et al. 1994) and the finite element method (Rahman
and Davies 1984, Koshiba 1992).*,† Approximate methods for com-
puting the effective index are also available, such as Marcatili’s
method (Marcatili 1969a), the effective index method (EIM) (Knox
and Toulios 1970), and perturbation methods (Chiang 1993). These
methods generally give good approximations for low index contrast
waveguides or for modes far from cutoff. Despite its approximate
nature, the EIM has found widespread use in the analysis of pla-
nar waveguides, even for high-index contrast waveguides, thanks

* The finite element method is typically formulated based on either Equation 1.9
for the electric field or Equation 1.22 for the magnetic field.
† Commercial software for computing the field distributions and effective indi-
ces of optical waveguide modes are also available, such as COMSOL, RSoft,
Optiwave, and Lumerical.
10 Optical Microring Resonators

to its simplicity and intuitive approach. Given the importance of


the method for waveguide analysis, we will briefly review the key
aspects of the EIM method below.

1.1.2 The EIM and solutions of the one-dimensional


slab waveguide
The basic idea of the EIM method (Knox and Toulios 1970) is the
successive approximations of a two-dimensional (2D) rectangular
waveguide by one-dimensional (1D) slab waveguides, which can be
separately analyzed. The procedure is illustrated in Figure 1.4 for
both a rib waveguide and a strip waveguide. In the first approxi-
mation, the vertical index profile in each of the core and cladding
regions (regions I and II) is replaced by the effective index of the
corresponding 1D, y-confined slab waveguides with y-dependent
index profiles. This procedure reduces the 2D waveguide to a 1D,
x-confined slab waveguide with an x-dependent effective index dis-
tribution neff(x). The effective index of the equivalent slab waveguide
(a) (b)
n3 y
n3 n3 n3 y n3
h
n1 n3 n1 t n3
n1 n1 t n1 x
n(y)
n2 n2 n2 n2 n2 n2
n2

y-slab I II I II II I II

neff,2 neff,1 neff,2 n3 neff,1 n3

neff(x) neff(x)
x-slab
neff,1 neff,1
neff,2 neff,2 n3 n3

x x

Figure 1.4 Successive approximations of a 2D waveguide by 1D slab


waveguides in the EIM: (a) rib waveguide and (b) strip waveguide. The
index profile n(y) of the y-confined slab waveguide in the core region
(region I) and the effective index distribution neff(x) of the x-confined slab
waveguide are also shown.
Elements of an Optical Microring Resonator 11

is then determined and taken as an approximation to the effective


index of the 2D rectangular waveguide.
The mathematical basis of the EIM method lies in the assump-
tion that the TE and TM semi-vectorial wave equations are separa-
ble. For example, for the quasi-TE mode, we assume that the solution
for the electric field Ex in Equation 1.19 has the form

Ex ( x , y ) = X( x)Y( y ). (1.29)

Substituting this solution into Equation 1.19 and dividing by X(x)


Y(y), we get

1 d  1 d 2  1 d 2Y
X dx  n2 dx
(
n X +
 Y dy
2)+ (n2 k 2 − βTE
2
) = 0, (1.30)

where n = n(x, y) is the index profile of the 2D waveguide. By adding


2
and subtracting the term neff ( x)k 2 to Equation 1.30, we can separate
it into two 1D wave equations (Okamoto 2000):

d 2Y
2
+  n 2 ( x , y ) − neff
2
( x) k 2Y( y ) = 0, (1.31)
dy

d  1 
 2
d 2
dx  neff ( x) dx
( )
neff ( x)X  +  neff
2
( x)k 2 − βTE
2
 X( x) = 0. (1.32)

We recognize Equation 1.31 as the TE wave equation for a y­ -confined


slab waveguide and Equation 1.32 is the TM wave equation for an
x-confined slab waveguide. We first solve Equation 1.31 in each of
the core and cladding regions (regions I and II) to obtain the lat-
eral effective index distribution neff(x). The TM effective index of the
equivalent x-confined slab waveguide is then determined by solv-
ing Equation 1.32. Alternatively, it is more convenient to determine
the TM effective index of the x-confined slab waveguide by solving
the wave equation in terms of the magnetic field Hy

d2 Hy
+  neff
2
( x)k 2 − β 2  H y ( x) = 0. (1.33)
dx 2 

In general, the error in the effective index value obtained by the


EIM method arises from two approximations. The first approxima-
tion is the use of the semi-vectorial wave equations to approximate
12 Optical Microring Resonators

the hybrid modes of the waveguide. The second approximation


comes from the fact that in order for the semi-vectorial equation
(1.19) or (1.20) to be separable, the index profile of the waveguide
must be decomposable in the form (Chiang 1996)

n2 ( x , y ) = nx2 ( x) + ny2 ( y ). (1.34)

The actual index functions nx(x) and ny(y) assumed by the EIM
method depend on the waveguide structure being analyzed.
The computation of the effective index of a 2D waveguide
by the EIM method reduces to the solution of two 1D slab wave-
guides. In fact, one of the appealing features of the EIM method
is that analytical solutions exist for the TE and TM modes of a 1D
slab waveguide. In Table 1.2, we summarize the field solutions and
­characteristic equations for the TE and TM modes of a general
asymmetric slab waveguide with width d and index distribution
n(x) shown in Figure 1.5.

Table 1.2 Summary of the Solutions and Characteristic Equations


for the TE and TM Modes in an Asymmetric Slab Waveguide
TE Modes TM Modes
Wave Equation
d 2 Ey d2 Hx
+ [n2 ( x)k 2 − β 2 ]Ey = 0 + [n2 ( x)k 2 − β 2 ]H x = 0
dx 2 dx 2
Field Solution
Ey ( x , z) = E0 ψ( x)e − jβz H x ( x , z) = H 0 ψ( x)e − jβz

cos(kx d/2 + θ)e − γ ( x − d/2) , x > d/2



ψ( x) = cos(kx x + θ), − d/2 ≤ x ≤ d/2
cos(k d/2 − θ)e α ( x + d/2) , x < − d/2
 x

Characteristic Equation
2kx d − ϕ 1 − ϕ 2 = 2mπ (m = 0, 1, 2, 3, …)
4θ = φ1 − φ2
φ1 = 2 tan−1(α/kx) ϕ 1 = 2 tan −1 (n12α/n22 kx )
φ2 = 2 tan−1(γ/kx) ϕ 2 = 2 tan −1 (n12 γ /n32 kx )
β 2 + kx2 = n12 k 2 , β 2 − α 2 = n22 k 2 , β 2 − γ 2 = n32 k 2
Elements of an Optical Microring Resonator 13
x

n3
d/2
n1
0 z
–d/2
n2

Figure 1.5 Schematic of a 1D asymmetric slab waveguide of width d, core


index n1, lower cladding index n2, and upper cladding index n3.

1.1.3 Waveguide dispersion


In general, the effective index of a dielectric waveguide depends
on the wavelength so that light at different frequencies propagates
at different velocities. This gives rise to dispersion effects such as
temporal broadening of a pulse propagating in the waveguide. In a
single-mode waveguide, the two main sources of this wavelength
dependence (also called intramodal dispersion) are material disper-
sion and waveguide dispersion. Material dispersion refers to the
dependence of the refractive indices of the core and cladding materi-
als on the wavelength. Material dispersion has its physical origin in
the dependence of the optical absorption of a material on frequency,
so that its permittivity also depends on the frequency through the
Kramers–Kronig relation. The dependence of the refractive index
of a material on the wavelength can be modeled by the Sellmeier
equation

Ai
n 2 (λ ) = 1 + ∑ 1 − ( λ /λ )
i i
2
. (1.35)

The coefficients Ai and λi for silica and crystalline silicon are given
in Table 1.3.
The second source of intramodal dispersion is waveguide dis-
persion, which is a structural effect and arises from the confine-
ment of light in the waveguide core. In general, strongly confined
waveguides exhibit higher waveguide dispersion due to stronger
interaction of the mode with the core boundaries. In addition, in
a multi-mode waveguide, there exists a third source of dispersion,
14 Optical Microring Resonators

Table 1.3 Coefficients of the Sellmeier Equation for SiO2 and Si


SiO2 Si
λ = 0.21–3.71 μm at 295 K λ = 1.1–5.6 μm at 295 K
(Malitson 1965) (Frey et al. 2006)
A1 0.6961663 10.67087
A2 0.4079426 −37.10820
A3 0.8974794
λ1 (μm) 0.0684043 0.3045744
λ2 (μm) 0.1162414 611.2222
λ3 (μm) 9.896161

called intermodal dispersion, which arises from the fact that dif-
ferent waveguide modes have different effective indices and thus
propagate at different phase velocities.
The parameter used to quantify the total dependence of
the effective index on wavelength is the group index, which is
defined as

dβ dn
ng = = neff − λ 0 eff . (1.36)
dk dλ

From the group index, we can calculate the group velocity, which is
the velocity at which a pulse with a frequency spectrum centered
around λ0 travels in the waveguide,

dω c
vg = = . (1.37)
dβ ng

The group delay experienced by a pulse after propagating a unit


distance in the waveguide is given by τg = 1/vg = dβ/dω. To express
the fact that the group delay is wavelength dependent, we write τg in
terms of a Taylor series expansion around the center wavelength λ0,

dτ g (∆λ )2 d 2 τ g
τ g (λ ) = τ g (λ 0 ) + ∆λ + + (1.38)
dλ 2 dλ 2

Defining the total chromatic dispersion of the waveguide as

dτ g d  ng  λ 0 d 2 neff
D= = = − , (1.39)
dλ dλ  c  c dλ 2
Elements of an Optical Microring Resonator 15

we can approximate Equation 1.38 by

τ g (λ ) ≈ τ g (λ 0 ) + D∆λ. (1.40)

For a pulse of spectral width Δλ, we obtain from Equation 1.40 the
spread in the group delay due to dispersion in the waveguide,

∆τ = τ g (λ ) − τ g (λ 0 ) ≈ D∆λ. (1.41)

Thus the chromatic dispersion D, typically quoted in units of ps/


nm/km, gives the delay spread per unit bandwidth per unit length
of the waveguide.
Figure 1.6 shows the plots of neff versus λ for the TE and TM
modes of an SOI waveguide consisting of a silicon core of 250 nm
thickness and 400 nm width embedded in SiO2. The group index
is calculated to be ng = 4.433 for the TE mode and ng = 4.349 for the
TM mode. The chromatic dispersion of the waveguide is D = −13.29
ns/nm/km for the TE mode and D = 0.972 ns/nm/km for the TM
mode. These values represent the total effects of material disper-
sion in the core and cladding materials as well as the structural

2.6

2.5
TE
2.4
Effective index, neff

2.3

2.2

2.1

2 TM

1.9 Full-vectorial
Semi-vectorial
1.8
EIM
1.7
1.45 1.5 1.55 1.6 1.65
Wavelength, λ (µm)

Figure 1.6 Wavelength dependence of the effective index of an SOI


waveguide consisting of a silicon core with cross-sectional dimensions
400 × 250 nm2 embedded in a SiO2 cladding.
16 Optical Microring Resonators

dispersion of the waveguide. We can also define the chromatic dis-


persion Dmat due to the material alone,

λ 0 d2 n
Dmat = − , (1.42)
c dλ 2

where n(λ) is the bulk index of the material. Near the 1.55 μm
wavelength, the material dispersion is −0.862 ns/nm/km for Si
and 21.9 ps/nm/km for SiO2. Comparing these values to the total
chromatic dispersion D of the waveguide reveals that waveguide
­dispersion plays a dominant role in silicon waveguides and indeed
in high-index contrast waveguides in general.
It is evident from Figure 1.6 that the effective index also depends
on the polarization. In general, the polarization dependence of the
effective index arises from the birefringence of the material as
well as the geometry of the waveguide. Isotropic materials such as
silicon and silica do not have material birefringence, although all
materials exhibit some birefringence under thermal or mechanical
stress. Thus in a silicon waveguide, the dependence of the effective
index on the polarization is a purely structural effect. The total bire-
fringence of a waveguide is defined as the difference between the
effective indices of two orthogonal polarization states, commonly
chosen to coincide with those of the TE and TM modes:

TM TE
B(ω) = neff (ω) − neff (ω). (1.43)

The frequency dependence of the birefringence gives rise to polar-


ization mode dispersion (PMD), which is defined as

1 dB
PMD = B(ω) − ω (ps/km). (1.44)
c dω

The PMD gives the differential time delay between the TE and TM
components of a pulse per unit propagating distance.

1.1.4 Propagation loss


There are three main sources of loss in an optical waveguide: o­ ptical
absorption in the core and cladding materials, electromagnetic scat-
tering, and radiation leakage. Optical absorption arises from various
electronic processes in the material such as atomic and molecular
Elements of an Optical Microring Resonator 17

vibrations in glass and polymers, and interband t­ransitions and


free carrier absorption (also known as intraband transitions) in semi-
conductors. Even when the waveguide is operated at a wavelength
far from an optical transition, there is still some residual absorption.
For example, for bulk crystalline Si, which has a bandgap of 1.1 μm,
accurate measurement of the optical absorption constant in a sam-
ple with a low impurity concentration of 2 × 1012 cm3 gives a value of
0.001 dB/cm at 1.55 μm wavelength (Steinlechner et al. 2013).
Loss due to electromagnetic scattering in an optical waveguide
is caused by two mechanisms: volume or Rayleigh scattering, and
surface roughness scattering. Rayleigh scattering refers to the scat-
tering of light by small fluctuations in the refractive index caused
by voids, defects, and contaminants in the material. Rayleigh
­scattering decreases with wavelength as λ−4, and is typically much
smaller than surface roughness scattering. The latter type of scat-
tering refers to the scattering of light due the roughness of wave-
guide surfaces caused by fabrication processes such as deposition
and etching. While the surface roughness due to deposition can be
controlled to less than 1 nm, the roughness of the waveguide side-
walls due to dry etching can be as large as a few nanometers. We
thus expect to have much larger scattering loss at the waveguide
sidewalls than at the top and bottom surfaces of the waveguide core.
In addition to the degree of roughness, surface scattering also
depends on the index contrast and how strongly light is confined in
the waveguide. Several methods have been developed for estimating
waveguide loss due to surface roughness scattering, ranging from
the simple model of Tien based on specular reflection (Tien 1971),
to more sophisticated models based on the Coupled Mode Theory
(CMT) (Marcuse 1969) which take into account the statistical distri-
bution of the roughness. However, since it is difficult to measure the
roughness profiles on the sidewalls of a waveguide, the usefulness of
these analyses is limited to providing broad ­estimates of the contri-
butions of surface roughness scattering to the total waveguide loss.
The third major source of waveguide loss is radiation leakage.
Radiation leakage arises in waveguide geometries which do not
have true eigenmode solutions. The two common types of radiation
loss in optical waveguides are bending loss and substrate leakage.
Bending loss occurs in curved waveguides and will be discussed
in more detail in Section 1.2. Substrate leakage refers to the leakage
of light from a waveguide into a high-index substrate. In theory,
the evanescent field of a waveguide extends indefinitely into the
18 Optical Microring Resonators

undercladding. The presence of a high-index substrate (such as sili-


con) causes evanescent coupling of light into the substrate which
radiates away as loss. Substrate leakage can be minimized by
increasing the thickness of the undercladding layer to provide suffi-
cient ­isolation of the waveguide core from the high-index substrate.

1.2 Optical Modes in Bent Dielectric Waveguides


In a bent dielectric waveguide, the optical mode is pushed toward
the outer edge of the waveguide as light propagates around the bend.
As the radius of curvature increases, a portion of the evanescent tail
of the mode begins to leak out in the form of radiation, resulting in
bending loss. A number of techniques have been used to analyze
the modes of curved optical waveguides and the associated bending
loss. These techniques range from approximate analytical methods
such as Marcatili’s method (Marcatili 1969b), the conformal map-
ping method (Heiblum and Harris 1975), to rigorous numerical solu-
tions of the wave equation in cylindrical coordinates (Rivera 1995,
Lui et al. 1998, Kakihara et al. 2006). In general, approximate analyti-
cal methods give adequately accurate results for curved waveguides
with low-index contrasts and large bending radii. For high-index
contrast and tightly bent waveguides, the modes become highly
hybridized and a full-vectorial numerical solution is required to
obtain an accurate analysis of the modal characteristics.
We begin in Section 1.2.1 with an approximate analysis of bent
waveguides by the conformal mapping method. Although strictly
valid only for 2D structures, the conformal mapping method pro-
vides an intuitive understanding of the propagation characteristics
and the mechanisms causing radiation loss in curved waveguides.
For microdisk and microring structures, analytical solutions for the
discrete resonant modes can be obtained by solving the 2D semi-
vectorial wave equation in polar coordinates. This is the subject
of Section 1.2.2. Finally, Section 1.2.3 will give a full-vectorial for-
mulation of the problem in the three-dimensional (3D) cylindrical
coordinate system (CCS) which is suitable for rigorous numerical
simulations of bent waveguides.

1.2.1 Conformal transformation of bent waveguides


The idea of the conformal mapping method is to apply a coordinate
transformation to the wave equation which will convert the curved
Elements of an Optical Microring Resonator 19

boundaries of the structure in the original (x, y) coordinates into


straight boundaries in the new (u, v) coordinates. Conformal map-
ping is based on the Cauchy–Riemann equations, which are valid
for domains in a 2D plane. To apply the method to a 3D curved
waveguide, we first reduce the waveguide to an equivalent 2D
structure in the x–y plane using the EIM, as shown in Figure 1.7.
The semi-vectorial equation governing wave propagation in the 2D
bent waveguide is then given by

∂2 F ∂2 F
2
+ 2 + n 2 ( x , y )k 2 F = 0, (1.45)
∂x ∂y

where n(x, y) is the effective index distribution and

Hz , quasi-TE mode,


F( x , y ) =  (1.46)
 Ez , quasi-TM mode.

The conformal transformation which converts circular bound-


aries in the x–y plane into straight boundaries in the u–v plane is
(Heiblum and Harris 1975)

Rref  x 2 + y 2   r 
u= ln  = Rref ln  , (1.47)
2  R 2
ref

  Rref 

 y
v = Rref tan −1   = Rref θ, (1.48)
 x

(a) z y (b) y

n2
n1
n2
r
n2 n1 n2
θ
0 0
n3 R1 R2 x R1 R2 x

Figure 1.7 Reduction of a 3D bent waveguide in (a) to an equivalent 2D


structure in (b) by the EIM. The index n1 of the core region is replaced by
the effective index n1 in (b).
20 Optical Microring Resonators

where Rref is a reference radius and (r, θ) are the polar coordinates of
a point in the x–y plane. Applying the above transformation to the
wave equation in Equation 1.45 gives

∂2 F ∂2 F
+ + n2 (u)k 2 F = 0, (1.49)
∂u 2 ∂v 2

where the index profile of the structure in the u–v plane is given by

n(u) = n( x , y )e u/Rref . (1.50)

Figure 1.8 shows the mapping of a microdisk and a microring


into straight-edge structures in the u–v plane. For the microdisk,
the reference radius Rref is chosen to be the microdisk radius R
whereas for the microring, it is taken to be the outer radius R 2. For
simplicity we have relabeled the effective index of the core as n1
(instead of n1). Note that for each structure, the step index profile
in the x–y plane is transformed into an index profile in the u–v
plane that is independent of v but varying in u with exponential
­dependence e u/Rref .
The nonlinear index profiles in Figure 1.8b and d of the trans-
formed waveguides reveal several important aspects of wave
propagation in microdisks and microrings. First, since the outer

(a) y (b) v (c) y (d) v


n2
n2 Inner n1
Inner n1(u) n2(u) n2(u) n1(u) n2(u)
caustic n1 caustic
0 R x 0 u 0 R1 R2 x u1 0 u
Ric Ric

n(r) n(u) n(r) n(u)


n1 n1
n1 n1
nv
nv
nv,2
n2 n2 n2
n2
n2

0 R r uic 0 urc ux u 0 R1 R 2 r u1 uic 0 urc ux u

Figure 1.8 Conformal mapping of a microdisk (a) and a microring (c) in


the x–y plane to straight-edge waveguides in the u–v plane, (b) and (d).
The index profiles n(r) in the x–y plane and n(u) in the u–v plane are also
shown.
Elements of an Optical Microring Resonator 21

cladding index exceeds the core index at the point u = ux, the
modes in a microdisk or microring are inherently leaky. A mode
with effective index nv = βv/k, as indicated in Figure 1.8b and d, is
confined within a “core” region uic < u < 0 for which nv < n1(u). At
the point u = 0 the wave is confined by total internal reflection (or
more precisely, by frustrated total internal reflection since there
is leakage into the high-index region beyond ux). The point uic,
which corresponds to the circle of radius R ic = (nv/n1)Rref in the
x–y plane, defines the inner caustic of the microdisk or microring
(shown in Figure 1.8a and c). As illustrated in Figure 1.8a, light
rays bouncing around the microdisk are tangential to the inner
caustic. These modes are called whispering gallery modes. For
a microring resonator, a mode behaves like a whispering gal-
lery mode if R ic > R1; otherwise, it is a regular waveguide mode
bounded by total internal reflection at both inner and outer
walls. For example, in Figure 1.8d, the mode with effective index
nv behaves like a whispering gallery mode while the mode with
effective index nv,2 is a regular waveguide mode. Note that since
the index distribution n1(u) in the core region increases toward
the outer waveguide wall (u = 0), we expect the field distributions
of the modes in a microdisk or microring to be skewed toward
the outer wall, with the skewness becoming more pronounced for
smaller bending radii.
At the point urc in the outer cladding region (u > 0) where nv
becomes smaller than n2(u), the mode becomes radiative. The point
urc defines the radiation caustic and corresponds to the radius
Rrc = (nv/n2)Rref at which the wave reaches the speed of light in the
cladding medium and radiates away. The radiation caustic can also
be determined by considering the microdisk in the x–y plane. As
the wave travels around the microdisk, the tangential speed (or the
speed in the θ direction) of the wave front must increase radially in
order to maintain a straight wave front. Using the relation nv = neff
exp(u/Rref) = neff(r/Rref), where neff is the effective index in the x–y
plane, we obtain for the tangential speed

c c  r 
v(r ) = = . (1.51)
neff nv  Rref 

The above equation shows that the tangential speed becomes equal
to the speed of light in the cladding medium, c/n2, at the ­radiation
caustic Rrc = (nv/n2)Rref.
22 Optical Microring Resonators

To obtain the modal field distribution and propagation constant


of the microring or microdisk waveguide, we can solve the wave
equation (1.49) in the u–v plane. The field solution has the form

F(u, v) = ψ(u)e − jγ v v , (1.52)

where ψ(u) is the radial field distribution and the propagation con-
stant γv = βv − jαv is complex due to the leaky nature of the mode. In
particular, the attenuation constant αv gives the bending loss of the
curved waveguide. Substituting (1.52) into Equation 1.49, we obtain
the eigenvalue problem

d2 ψ
+ n2 (u)k 2 ψ = γ 2v ψ , (1.53)
du2

which can be solved for the propagation constant γv and the field
distribution ψ(u). Since the structure has a graded index profile n(u),
Equation 1.53 does not have analytical solutions and must be solved
using a numerical technique such as the finite difference method.
Approximate analytical solutions have also been obtained using
the WKB (Wentzel–Kramers–Brillouin) method (Berglund and
Gopinath 2000) or by linearizing the index profile to obtain an Airy-
type equation (Chin and Ho 1998). The solution obtained in the u–v
plane is then converted back to the x–y (or r–θ) plane to give

F(r , θ) = ψ(Rref ln(r/Rref ))e − jγ v Rref θ . (1.54)

It is evident from the above solution that γv gives the propagation


constant along the arc length of the curved waveguide at the radial
distance Rref from the center. Since the effective index in the x–y
plane is given by neff = nv(Reff/r), we obtain the expression for the
propagation constant in the x–y plane as

Rref R
γ = neff k = γ v = ref (β v − jα v ), (1.55)
r r

which shows that γ, and hence neff, vary as 1/r. It is typical to define
the effective propagation constant of a microdisk or microring as
the value of γ at the effective radius Reff,

Rref
γ eff = (β v − jα v ), (1.56)
Reff
Elements of an Optical Microring Resonator 23

where Reff is the radial distance to the centroid of the intensity


­distribution of the mode (Rowland and Love 1993):
2


Rrc
ψ(r ) dr
Reff = 0
. (1.57)
1 2


Rrc

r ψ ( r ) dr
0

For microrings, the effective radius is very close to the average


radius, Ravg = (R1 + R2)/2, and in most cases can be approximated
by this value.

1.2.2 Resonant modes in microdisks and microrings


In the previous section we have assumed that the waveguide is
unbounded in the propagation direction (i.e., the v direction in the
u–v plane or the θ direction in the r–θ plane). As a result, the propa-
gation constant βv is a continuous function of frequency. In a micro-
disk or microring resonator, the imposition of the periodic boundary
condition in θ (or v) gives rise to discrete eigenmodes at discrete
frequencies, which are the resonant modes of the structure. For 2D
microdisks and microrings, we can obtain analytical solutions for
the field distributions of the eigenmodes and their resonant frequen-
cies by solving the semi-vectorial wave equation in polar coordinates
subject to the periodic boundary condition.
We consider first the solution for a microdisk with radius R. As
in the previous section, we assume that the 3D microdisk structure
has been reduced to an equivalent 2D structure using the EIM, with
the effective index distribution in the r–θ plane given by
n1 , r ≤ R,
n(r ) =  (1.58)
 n2 , r > R.
In polar coordinates, the equation governing wave propagation in
the microdisk is
∂ 2 F 1 ∂F 1 ∂ 2 F
∇ 2 F + n2 k 2 F = + + + n2 (r )k 2 F = 0, (1.59)
∂r 2 r ∂r r 2 ∂θ 2
where k = ω/c and

 H z , quasi-TE mode,
F( r , θ ) = 
Ez , quasi-TM mode.
24 Optical Microring Resonators

Equation 1.59 can be solved using the method of separation of vari-


ables. Letting F(r,θ) = ψ(r)Θ(θ), we separate the equation to get

d 2Θ
+ βθ2Θ = 0, (1.60)
dθ 2

d 2 ψ 1 dψ  2 2 βθ2 
+ + n ( r )k − ψ = 0, (1.61)
dr 2 r dr  r 2 

where βθ2 is the constant of separation. The solution of Equation 1.60


subject to the periodic boundary condition Θ(θ) = Θ(θ + 2π) is

Θ m (θ) = e ± jmθ , m = 0, 1, 2, 3, … (1.62)

with the separation constant given by βθ2 = m2. The integer m denotes
the azimuthal number of the resonant mode. Substituting βθ = m
into Equation 1.61, we get

d 2 ψ 1 dψ  2 m2 
2
+ +  n (r )k 2 − 2  ψ = 0, (1.63)
dr r dr  r 

which is the Bessel equation of order m. Assuming waves in the


cladding region propagate outward in the radial direction, we write
the solution to Equation 1.63 as

C1 J m (n1kr ), r ≤ R,
ψ(r ) =  ( 2) (1.64)
C2 H m (n2 kr ), r > R,

where Jm and H m( 2) are the Bessel function and Hankel function of


the second kind, respectively, of order m. Since ψ must be continu-
ous at the dielectric interface at r = R, we obtain the relation for the
amplitude coefficients,

J m (n1kR)
C2 = C1 . (1.65)
H m( 2) (n2 kR)

In addition, we also require that Eθ ∝ (1/n2)dψ/dr be continuous at


r = R for the TE mode, and Hθ ∝ dψ/dr be continuous at r = R for
the TM mode. Enforcing the above field continuity conditions and
Elements of an Optical Microring Resonator 25

making use of Equation 1.65, we obtain the characteristic equation


for the microdisk*

J m′ (n1kR) sH m( 2)′ (n2 kR)


= , (1.66)
J m (n1kR) H m( 2) (n2 kR)

where s = n1/n2 for TE and s = n2/n1 for TM. Solution of the above
equation gives the eigenvalues kmn, where n = 1, 2, 3, … represents
the radial mode number. Since the modes in a microdisk are leaky,
kmn is a complex number. The characteristic frequency of mode (m, n)
is thus also complex, which can be expressed as
ω mn = ω mn
′ + jω mn
′′ = ck mn . (1.67)

Using the above expression, we can write the time-dependent solu-


tion for the resonant mode (m, n) in the microdisk as

F(r , θ, t) = ψ mn (r )e ± jmθ e jω mnt = ψ mn (r )e j(ω mn


′ t ± mθ ) − ω mn
e ′′ t , (1.68)

where the plus and minus signs correspond to the counterclockwise


and clockwise propagating modes, respectively. In Equation 1.68, ω mn ′
is the resonant frequency of mode (m, n) and ω mn ′′ gives the temporal
rate of field decay in the cavity due to bending loss. The angular veloc-
ity of the phase front of mode (m, n) is vθ , mn = ω mn
′ /m. We may define
the effective index of mode (m, n) from the tangential velocity (rvθ,mn) of
the wave at the effective radial distance Reff from the microdisk center,
c mλ mn
neff , mn = = , (1.69)
Reff vθ , mn 2πReff

where λ mn = 2πc/ω ′mn is the resonant wavelength of mode (m, n).


The above equation can be recast in the form
mλ mn
= 2πReff , (1.70)
neff , mn

* At nonresonant frequencies, the azimuthal equation (1.60) does not have to sat-
isfy the periodic boundary condition so it must be treated as an open bound-
ary problem. In this case the separation constant βθ = ν is an unknown complex
number and we must replace Jm and H m( 2) by Jν and H (ν2), respectively, in the g
­ eneral
solution for ψ in (1.64). The characteristic equation (1.66) is then an ­equation in
terms of the complex order ν of the Bessel and Hankel functions. Solution of
the characteristic equation involving Bessel and Hankel functions of complex
orders has been investigated for bent waveguides in Hiremath et al. (2005).
26 Optical Microring Resonators

which indicates that at resonance, the effective microdisk circum-


ference (2πReff) must be equal to an integer multiple of the guided
wavelength.
The attenuation constant of the mode due to bending loss can be
obtained by converting the temporal decay factor in Equation 1.68
to an angular attenuation factor:

exp(−ω ′′mn t) = exp ( −ω ′′mn θ/vθ , mn ) = exp ( − mω mn ′ ).


′′ θ/ω mn (1.71)

From the above expression, we obtain the power attenuation con-


stant αmn along the arc length at a radial distance Reff from the
microdisk center to be

2mω mn ′′
α mn = . (1.72)
ω ′mn Reff

Another parameter of interest is the quality factor (Q factor) of the


microdisk resonator, which is defined as 2π times the ratio of the
time-averaged energy stored to the energy loss per cycle (Jackson
1999):

Average stored energy Average stored energy


Q = 2π = ω ′mn .
Power loss × Optical period Power loss
(1.73)

Since the instantaneous energy density in the microdisk is propor-
tional to the square of the real field in Equation 1.68, we can express
the total stored energy in the resonator at time t as

′ t)e −2ω mn
U(t) = U 0 cos 2 (ω mn ′′ t
, (1.74)

where U0 is the initial stored energy. Assuming that the rate of


energy decay is much smaller than the angular frequency of the
optical field (2ω mn
′′ << ω mn
′ ), the stored energy in the microdisk
­averaged over one optical cycle is

1
U = U 0 e −2ω mn
′′ t
. (1.75)
2

Since the power loss is just the rate of energy decay,

dU
PL = − = 2ω ′′mn U , (1.76)
dt
Elements of an Optical Microring Resonator 27

we obtain the quality factor associated with the resonant mode


(m, n) as

U ω′
Q = ω ′mn = mn . (1.77)
PL 2ω ′′mn

If we define the cavity lifetime (or photon lifetime) τmn as the time
it takes for the stored energy to decay to 1/e of its initial value, we
have from Equations 1.75 and 1.77 that τ mn = 1/2ω ′′mn = Q/ω ′mn. Thus
the Q factor (divided by 2π) gives the cavity lifetime in terms of the
number of optical cycles.
Solutions of the resonant modes in a microring resonator can be
obtained in a similar manner as for the microdisk. The only differ-
ence lies in the solution of Equation 1.63 for the radial field distribu-
tion. For a 2D microring with inner radius R1, outer radius R2, and
effective index distribution

n1 , R1 ≤ r ≤ R2 ,
n(r ) =  (1.78)
n2 , r < R1 and r > R2 ,

the solution for the radial field distribution has the form

C1I m (n2 kr ), r < R1 ,



ψ(r ) = C2 J m (n1kr ) + C3 Ym (n1kr ), R1 ≤ r ≤ R2 , (1.79)
 C H ( 2) (n kr ), r > R2 ,
 4 m 2

where Ym and Im are, respectively, the Neumann function and the


modified Bessel function of the first kind of order m. Again the
Hankel function is chosen as the solution in the outer cladding
region, r > R2, to allow for radiation leakage. By matching the tan-
gential fields at the inner and outer radii, we obtain the following
characteristic equation for the microring:

sI m′ (n2 kR1 ) J m (n1kR1 ) − I m (n2 kR1 ) J m′ (n1kR1 )


sH m( 2)′ (n2 kR2 ) J m (n1kR2 ) − H m( 2) (n2 kR2 ) J m′ (n1kR2 )
sI m′ (n2 kR1 )Ym (n1kR1 ) − I m (n2 kR1 )Ym′ (n1kR1 ) (1.80)
= ,
sH m( 2)′ (n2 kR2 )Ym (n1kR2 ) − H m( 2) (n2 kR2 )Ym′ (n1kR2 )

where s = n1/n2 for TE modes and s = n2/n1 for TM modes.


Numerical results for the resonant modes in silicon microrings
and microdisks obtained using the above analysis are shown in
28 Optical Microring Resonators
(a) (b)
1 1

Normalized electric field amplitude, Ez


Normalized electric field amplitude, Er
TE TM
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3
Radial distance, r (µm) Radial distance, r (µm)

Figure 1.9 Transverse electric field amplitude of the resonant mode


(m, n) = (23, 1) in a silicon microdisk with 2 μm radius (dashed line) and a
silicon microring with R1 = 1.7 μm and R 2 = 2 μm (solid line): (a) Er field of
the TE mode, (b) Ez field of the TM mode.

Figures 1.9 and 1.10. In all the plots we assume a Si waveguide core
of thickness of 250 nm embedded in a SiO2 cladding. Figure 1.9a
and b show the transverse electric field distributions of the TE and
TM resonant modes with m = 23, n = 1, in a microdisk with a 2 μm
radius and a microring with R1 = 1.7 μm and R2 = 2 μm. A slight
skewing of the modes toward the outer edge is noticeable, espe-
cially for the TM mode. Figure 1.10a and b show the effective indi-
ces of the TE and TM resonant modes as functions of the resonant
wavelength for microrings with a fixed average radius of 2 μm and
microring width varying from 250 nm to 1 μm. In both plots, lines
of constant azimuthal mode number m and constant ring width are
also shown. We observe that the effective index of the microring
exhibits strong dispersion, especially for small ring widths. As the
ring width increases, the microring behaves more like a whispering
gallery mode resonator so its dispersion characteristic approaches
that of a microdisk, whose effective index is shown by the bold
dark line in both plots. Figure 10c and d show the dependence of
the Q factor of the microring on the resonant wavelength and the
ring width. The bold dark line in each plot indicates the Q factor
of the microdisk, which serves as the limiting case for microrings
with large ring widths. From these plots we observe that for a fixed
ring width, the Q factor decreases with wavelength since the mode
becomes less confined, leading to higher radiation leakage due to
bending.
Elements of an Optical Microring Resonator 29
(a) (b)
m = 15 Increasing m TE Increasing m TM
3.3 m = 15
3.2
3.2
3 3.1
Effective index, neff

Effective index, neff


3
2.8 m = 25
2.9 m = 25
2.6 2.8
2.4 2.7
2.6
2.2 2.5 Increasing
Increasing ring width
2 2.4
250 nm ring width 250 nm
2.3
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 1.4 1.6 1.8 2 2.2 2.4 2.6
Resonant wavelength, λ (µm) Resonant wavelength, λ (µm)

(c) (d) 1010


1011
m = 15 TE m = 15 TM
1010 Increasing m 109
Increasing m
109
108
Quality factor, Q
Quality factor, Q

108
107 107
106 106
105 m = 25
105 m = 25
104
104 Increasing
103 Increasing
ring width 250 nm ring width
250 nm 103
102
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 1.4 1.6 1.8 2 2.2 2.4 2.6
Resonant wavelength, λ (µm) Resonant wavelength, λ (µm)

Figure 1.10 Effective indices of the resonant modes of a silicon micror-


ing with core thickness of 250 nm, average radius of 2 μm, and ring width
varying from 250 nm to 1 μm in steps of 50 nm: (a) TE mode, (b) TM mode.
Quality factors of (c) TE and (d) TM resonant modes in the microrings.
The bold dark lines in all the plots are the results for a silicon microdisk
with 2.5 μm radius.

Figure 1.11 shows the dependence of the bending loss of a


s­ ilicon microring on the average ring radius for TE and TM reso-
nant modes near the 1.55 μm wavelength. The microring width is
fixed at 400 nm. The bending loss is seen to exhibit an exponential
dependence on the bending radius. This exponential dependence
has also been predicted by Marcuse’s approximate analytical for-
mula for loss in a bent slab waveguide (Marcuse 1972). Due to the
high-index contrast of the SOI material system, the bending loss of
silicon microrings is seen to remain low for very tight bends with
radii down to 2.5 μm. For comparison we also show the bending
loss of SiN microrings with SiO2 cladding. The SiN core thickness
is assumed to be 400 nm and the ring width is 800 nm. Since the
SiN/SiO2 material system has a lower index contrast, the bending
30 Optical Microring Resonators

104

TE
102
TM

Bending loss (dB/µm)


100

10–2
SOI SiN

10–4

10–6

1 2 3 4 5 6 7 8 9 10 11 12
Average bend radius, Ravg (µm)

Figure 1.11 Bending loss as a function of the average bend radius of SOI
and SiN microrings. The dimensions of the silicon ring waveguide are
250 × 400 nm (height × width), while those of the SiN ring waveguide are
400 × 800 nm. The bending losses are computed for resonant modes near
the 1.55 μm wavelength.

loss lines for SiN microrings are shifted to larger radii compared
to those of SOI microrings, although compact bends with radii less
than 10 μm can still be achieved with low loss in SiN.

1.2.3 Full-vectorial analysis of bent waveguides


The approximate analytical solutions in the previous sections are
useful for gaining an intuitive understanding of the characteristics
and behavior of wave propagation in bent waveguides. For an accu-
rate determination of the modes and bending loss in a 3D struc-
ture, however, we must resort to a full-vectorial analysis due to the
high degree of hybridization of the modes. This effect is especially
pronounced in tight bends in high-index contrast materials such as
SOI, where low loss waveguide bends and microrings with radii
approaching 1 μm have been demonstrated (Vlasov and McNab
2004, Xu et al. 2008, Prabhu et al. 2010). A full vectorial analysis
of curved waveguides requires a numerical solution of the vector
wave equation, which is typically formulated in a local cylindrical
coordinate system (CCS) ­centered about the waveguide core.
We consider a bent waveguide with inner radius R1, outer
radius R2, and cross-sectional index distribution n(r, z) as shown
Elements of an Optical Microring Resonator 31
(a) z (b)
y

n1 n2
θ n2 n1 n2
0 0 x
n3 R1 R2 r n3
s

Figure 1.12 Waveguide bend in (a) the global cylindrical coordinate


­system (r, θ, z) and (b) the local cylindrical coordinate system (x, y, s).

in Figure 1.12a. In the CCS, we decompose the electric field in


the waveguide into a transverse and angular component as follows:

E(r , θ, z) = E t (r , θ, z) + Eθ (r , θ, z)θ , (1.81)

where E t = Er rˆ + Ez zˆ and θ̂ is the unit vector in the angular direc-


tion. Substituting the above expression into the wave equation in
Equation 1.9,

∇ 2E + n2 k 2E = ∇(∇ ⋅ E ), (1.82)

we obtain the following equation for the transverse field


c­ omponent Et:*

1 ∂ 2E t 1 
∇ t 2E t + 2 2
+ n2 k 2E t = ∇ t (∇ t ⋅ E t ) − ∇ t  2 ∇ t ⋅ (n 2E t ) , (1.83)
r ∂θ n 

where

∂ ∂ 1 ∂  ∂  ∂2
∇t = rˆ + zˆ and ∇ 2t = r  + .
∂r ∂z r ∂r  ∂r  ∂z 2

It is more convenient to solve Equation 1.83 in a local CCS


(x, y, s) centered about the waveguide core, as shown in Figure 1.12b.
The transformations from the global CCS (r, θ, z) to the local CCS are
given by (Cheng et al. 1990)
* Equation 1.83 is similar to Equation 1.13 except that it is expressed in cylindrical
coordinates.
32 Optical Microring Resonators

r → x + R, θ → s/R, z → y , (1.84)

where R = (R1 + R2)/2 is the average bending radius. We also rename


the transverse field components Er → Ex, Ez → Ey and assume an
e−jβs dependence in the propagation direction. The transverse field
can now be written as

E t = Ex ( x , y )x + Ey ( x , y )y  e − jβs . (1.85)

Applying the above coordinate transformations to Equation 1.83,


we obtain a system of equations for the transverse electric field
components which can be put in matrix form as follows (Lui et al.
1998, Feng et al. 2002):

 Pxx Pxy   Ex   Ex 
P    = β2   , (1.86)
 yx Pyy  Ey  Ey 

where
2
1 ∂ ρ ∂  ∂ Ex
Pxx Ex =
ρ2 ∂x  n2 ∂x
ρn 2
E (
x 

+
∂ y 2 )
+ n2 k 2 Ex , (1.87)

1 ∂  ρ2 ∂ 2  ∂ Ey
2
Pxy Ey =  n Ey  (
− , ) (1.88)
ρ2 ∂x  n 2 ∂y  ∂y ∂x

1 ∂  ∂Ey  ∂  1 ∂ 2 
Pyy Ey = ρ + 
ρ ∂x  ∂x  ∂y  n2 ∂y
(
n Ey  + n 2 k 2 Ey ,) (1.89)

∂  1 ∂  1 ∂  ∂Ex 
Pyx Ex =  (
ρn 2 Ex  − ρ ), (1.90)
∂y  ρn ∂x  ρ ∂x  ∂y 
2

and ρ = 1 + x/R. In the limit R → ∞, the above equations reduce to


Equations 1.15 through 1.18 for a straight waveguide.
We can also formulate a similar set of equations in terms of the
transverse magnetic field. From the wave equation (1.22) for the
magnetic field,

1
∇ 2H + n2 k 2H = − 2
∇n2 × (∇ × H ), (1.91)
n
Elements of an Optical Microring Resonator 33

we make the substitution H(r , θ, z) = Ht (r , θ, z) + Hθ (r , θ, z)θ to


obtain the following equation for the transverse field component in
the global CCS (Lui et al. 1998):

1 ∂ 2 Ht 1
∇ 2t Ht + 2 2
+ n2 k 2Ht = − 2 (∇ t n2 ) × (∇ t × Ht ). (1.92)
r ∂θ n

Writing the transverse magnetic field in the local CCS as

Ht =  H x ( x , y )x + H y ( x , y )y  e − jβs , (1.93)

and applying the coordinate transformations in Equation 1.84 to


Equation 1.92, we obtain the matrix equation (Lui et al. 1998, Xiao
et al. 2009)

Qxx Qxy   H x   Hx 
Q    = β2   , (1.94)
 yx Qyy   H y   Hy 

where

∂  ∂(ρH x )  2 2 ∂  1 ∂H x 
Qxx H x = ρ +ρ n + ρ2 n 2 k 2 H x , (1.95)
∂x  ∂x  ∂y  n2 ∂y 

∂  2 ∂H y  ∂  1 ∂H y 
Qxy H y =  ρ  − ρ2 n 2 , (1.96)
∂x  ∂x  ∂y  n2 ∂x 

2
∂  ρ ∂H y 
2 2 ∂ Hy
Qyy H y = ρn   +ρ + ρ2 n 2 k 2 H y , (1.97)
∂x  n ∂x 
2
∂y 2

∂ 2 Hx ∂  1 ∂H x 
Qyx H x = ρ2 − ρ2 n 2 , (1.98)
∂x ∂y ∂x  n2 ∂y 

and ρ = 1 + x/R. The above equations simplify to Equations 1.25


through 1.28 for a straight waveguide in the limit R → ∞.
Equations 1.86 and 1.94 can be solved using the finite difference
method (Lui et al. 1998) for the transverse field distributions and the
complex propagation constant β. Alternatively, one can also solve
Equation 1.82 or 1.91 directly in cylindrical coordinates using the
34 Optical Microring Resonators

finite element method (Kakihara et al. 2006). Due to the leaky nature
of the modes, care should be taken to apply appropriate absorbing
boundary conditions (ABCs) at the boundaries of the computational
domain to properly absorb the radiating waves. In particular, the
perfectly matched layer ABCs have been shown to be effective in
both the finite difference and finite element solutions (Feng et al.
2002, Kakihara et al. 2006).
Figure 1.13a and b show the transverse magnetic field distribu-
tions in the local CCS of the lowest quasi-TM mode of an SOI bent
waveguide (Prabhu et al. 2010). The waveguide dimensions are
340 × 300 nm2 (height × width) and the average bending radius is
1 μm. The skewing of the field distributions toward the outer radius
(in the positive x ­direction) is apparent. Figure 1.13c and d show the
roundtrip loss and the Q factor at the 1.55 μm wavelength as functions

(a) 1 (b) 1
air air
Hx Hy
0.5 0.5
y (µm)

y (µm)

0 0

–0.5 –0.5
SiO2 SiO2
–1 –1
–1 –0.5 0 0.5 1 –1 –0.5 0 0.5 1
x (µm) x (µm)
(c) 101 (d) 108
TE
107
100 TM
Roundtrip loss (dB)

106
Intrinsic Q factor

250 × 450 nm2


10–1 105
340 × 300 nm2 250 × 450 nm2
104
10–2
340 × 300 nm2
103
TE
10–3
TM 102

10–4 101
0.5 1 1.5 2 0.5 1 1.5 2
Average bend radius, Ravg (µm) Average bend radius, Ravg (µm)

Figure 1.13 (a) and (b) Transverse magnetic field components Hx and Hy
of the quasi-TM mode of an SOI curved waveguide with average radius of
1.0 μm and waveguide dimensions of 340 nm × 300 nm (height × width).
Dependence of the roundtrip bending loss (c) and the intrinsic Q factor (d)
of SOI curved waveguides on the bending radius at 1.55 μm wavelength
for two sets of waveguide dimensions, 340 × 300 nm2 and 250 × 450 nm2.
(From Prabhu, A.M., et al., 2010, IEEE Photonics J., 2(3): 436–444.)
Elements of an Optical Microring Resonator 35

of the ­bending radius for SOI microrings with two waveguide


aspect ratios (height × width): 340 × 300 nm2 and 250 × 450 nm2. We
observe that similar to the trend shown in Figure 1.11, the bending
loss decreases exponentially with increasing radius for both the TE
and TM polarizations. Note also the strong dependence of the bend-
ing loss and the Q factor on the polarization. For example, for a 2-μm
radius microring with aspect ratio 250 × 450 nm2, the bending losses
and Q factors of the TE and TM modes differ by as much as 4 orders
of magnitude. Also noteworthy from these plots is the fact that
fairly high intrinsic Q values can still be achieved with extremely
small microring resonators. For example, for the 1-μm radius
­microring with aspect ratio 250 × 450 nm2, the intrinsic Q ­factor
associated with bending loss exceeds 20,000 for the TE mode.

1.3 Coupling of Waveguide Modes in Space


When two dielectric waveguides are brought into close proximity
of each other, the evanescent tail of the modal field distribution of
each waveguide interacts with the other waveguide core, resulting
in coupling and power exchange between the two waveguides. This
evanescent coupling of power is the primary means by which light
is coupled into a microring resonator from an external waveguide,
or between two microring resonators. Evanescent coupling between
two dielectric waveguides is analyzed by means of the Coupled
Mode Theory (CMT), which describes the coupling of modes of
the two waveguides as they propagate in space. The CMT equa-
tions can be formulated from Maxwell’s equations using a number
of approaches such as perturbation theory (Yariv 1973), variational
approach (Haus et al. 1987), and reciprocity (Chuang 1987a,b). In
this section we will adopt the reciprocity approach to formulate the
equations for the coupling of waveguide modes in space and derive
the solution for a pair of evanescently coupled waveguides.

1.3.1 The coupled mode equations


We begin by deriving a general reciprocity relation for the electric
and magnetic fields in two media (or structures) described by the
z-invariant permittivity functions εa(x, y) and εb(x, y) (Chuang 1987a).
The fields E a, Ha and Eb, Hb in media a and b satisfy the Maxwell’s
equations

∇ × E a = − jωµ 0 Ha , (1.99)
36 Optical Microring Resonators

∇ × Ha = jωε a ( x , y )E a , (1.100)

∇ × Eb = − jωµ 0 Hb , (1.101)

∇ × Hb = jωε b ( x , y )Eb . (1.102)

By expanding the expression ∇·(E a × Hb − Eb × Ha) and making use


of the above equations, we get

∇ ⋅ (E a × Hb − Eb × Ha ) = − jω(ε b − ε a )E a ⋅ Eb . (1.103)

Taking the integral of Equation 1.103 over the cross-sectional


area in the x–y plane, we obtain with the help of the divergence
theorem,


∂z ∫ ∫
(E a × Hb − Eb × Ha ) ⋅ z dxdy = − jω (ε b − ε a )E a ⋅ Eb dxdy. (1.104)

The above reciprocity relation holds for any two sets of fields Ea, Ha
and Eb, Hb satisfying Maxwell’s equations in any two media εa(x, y)
and εb(x, y). Physically, it describes the interaction between the two
sets of fields through the polarization current Jj = jω(εi − εj)Ej induced
in each medium i, j = {a, b}.
We use the reciprocity relation in Equation 1.104 to formulate the
CMT equations for a system of two coupled parallel waveguides sepa-
rated by a gap s, as shown in Figure 1.14c. The permittivity functions
of the isolated waveguides are given by ε1(x, y) and ε2(x, y), as shown in
Figure 1.14a and b, and that of the coupled w­ aveguide pair is ε(x, y). Let
us choose medium a to represent the waveguide pair, εa = ε(x, y), and
medium b to be waveguide 1 in ­isolation, εb = ε1(x, y). In the coupled

(a) (b) (c)


ε1(x, y) εclad ε2(x, y) εclad ε(x, y) εclad

εc1 εc2 εc1 s εc2

εsub εsub εsub

Figure 1.14 (a) and (b) Cross sections of isolated waveguides 1 and 2
described by permittivity functions ε1(x, y) and ε2(x, y). (c) Coupled wave-
guide pair with permittivity function ε(x, y).
Elements of an Optical Microring Resonator 37

waveguides, we approximate the forward-­propagating fields in terms


of a linear superposition of the modes of waveguides 1 and 2,*

E a = a1 ( z)e(1) ( x , y ) + a2 ( z)e( 2) ( x , y ), (1.105)

Ha = a1 ( z)h(1) ( x , y ) + a2 ( z)h( 2) ( x , y ), (1.106)

where e(1), h(1) and e(2), h(2) are the electric and magnetic field distri-
butions of the modes of the isolated waveguides 1 and 2 with the
power normalization


1 (i)
e × h( i ) ⋅ z dx dy = 1, i = {1, 2}.
2

In subsequent derivations, it is convenient to decompose the modal


fields into a transverse and longitudinal component as

e(1) ( x , y ) = e(t1) + z e(z1) , (1.107)

h(1) ( x , y ) = h(t1) + z hz(1) , (1.108)

with similar expressions for e(2), h(2). In medium b, we take the fields
Eb, Hb to be the backward-propagating fields in waveguide 1:

Eb = e(1) ( x , y )e jβ1z = (e(t1) − z e(z1) )e jβ1z , (1.109)

* To be more accurate, we should express only the transverse components of


E a and H a by a linear superposition of the transverse fields of the modes of
­waveguides 1 and 2:
E(ta ) = a1 ( z)e(t1) ( x , y ) + a2 ( z)e(t2) ( x , y ),

H(ta ) = a1 ( z)h(t1) ( x , y ) + a2 ( z)h(t2) ( x , y ).


The longitudinal field components of E a and H a are then derived from
Maxwell’s equations to give (Chuang 1987a)
ε1 (1) ε
Ez( a ) = a1 ( z) ez ( x , y ) + a2 ( z) 2 e(z2) ( x , y ),
ε ε
H z( a ) = a1 ( z)hz(1) ( x , y ) + a2 ( z)hz( 2) ( x , y ).

The above field expressions lead to a slight modification to Equation 1.114 for
the coupling term:
ω ε i (i) ( j) 
∫ ∫ (ε − ε ) e

Ki, j = j
(i)
t ⋅ e(t j ) − ez ez  dx dy.
4 ε 
38 Optical Microring Resonators

Hb = h(1) ( x , y )e jβ1z = (−h(t1) + z hz(1) )e jβ1z , (1.110)

where β1 is the propagation constant of waveguide 1. By substituting


the two sets of fields in Equations 1.105, 1.106, and 1.109, 1.110 into
the reciprocity relation (1.104), we obtain after some simplification

da1 da
+ C 2 = j(K11 + β1 )a1 ( z) + j(K 21 + Cβ1 )a2 ( z), (1.111)
dz dz

where

C12 + C21
C= , (1.112)
2


1 (i)
Ci , j = et × h(t j ) ⋅ z dx dy, {i , j} = {1, 2}, (1.113)
2

ω
Ki, j =
4 ∫
(ε − ε j )e( i ) ⋅ e( j ) dx dy
(1.114)
ω
=
4 ∫
(ε − ε j )[e(ti ) ⋅ e(t j ) − ez( i ) ez( j ) ]dx dy , {i , j} = {1, 2}.

In a similar manner, if we choose medium a to be the wave-


guide pair, εa = ε(x, y), and medium b to be waveguide 2, εb = ε2(x, y),
we obtain a second equation in terms of the mode amplitudes a1
and a2,

da1 da2
C + = j(K12 + Cβ 2 )a1 ( z) + j(K 22 + β 2 )a2 ( z), (1.115)
dz dz

where β2 is the propagation constant of waveguide 2. The coupling


terms K12 and K 22 are given by Equation 1.114. Equations 1.111 and
1.115 can be combined as

d  a1 ( z)   a1 ( z) 
C   = − j(K + BC)  , (1.116)
dz  a2 ( z)  a2 ( z ) 

where

1 C  K11 K12  β1 0


C=  ,K=   ,B= . (1.117)
C 1  K 21 K 22  0 β 2 
Elements of an Optical Microring Resonator 39

Upon multiplying Equation 1.116 by C−1, we obtain the coupled


mode equation

da
= − jMa , (1.118)
dz

where a(z) = [a1(z), a2(z)]T and M = C−1(K + BC). The elements of the
coupling matrix M are

 γ1 k12 
M= , (1.119)
 k 21 γ 2 

with

γ 1 = β1 + (K11 − CK 21 )/∆ C , (1.120)

γ 2 = β 2 + (K 22 − CK12 )/∆ C , (1.121)

k12 = (K12 − CK 22 )/∆ C , (1.122)

k 21 = (K 21 − CK11 )/∆ C , (1.123)

and ΔC = 1 − C2. The constants γ1 and γ2 are the self-coupling or


phase constant of each waveguide, while k12 and k21 give the mutual
coupling strengths between waveguide 1 and waveguide 2. If the
two waveguides are identical, then β1 = β2, K11 = K22, and K12 = K21.
For this symmetric case we have γ1 = γ2 and k12 = k21 = kc.
Figure 1.15a and b show the coupling strength kc as a func-
tion of the coupling gap between two identical SOI waveguides
for the quasi-TE and quasi-TM modes. The waveguide thickness
is 250 nm and three different values of the waveguide width are
considered: 300, 400, and 500 nm. The plots show that the coupling
strength decreases exponentially with increasing coupling gap.
Also note that the coupling strength decreases with increasing
waveguide width. This is due to the fact that the modes are more
tightly confined in the cores and thus have weaker interaction with
each other.

1.3.2 Solution of the coupled mode equations


To solve the coupled mode equation (1.118) for a pair of coupled
waveguides, we diagonalize the coupling matrix M in the form
40 Optical Microring Resonators
(a) (b)
1.4 1
TE TM

Coupling strength, kc (µm–1)


Coupling strength, kc (µm–1)
1.2
0.8
1
0.8 300 nm 0.6 300 nm
0.6 0.4
400 nm 400 nm
0.4
0.2
0.2
500 nm 500 nm
0 0
0 100 200 300 400 500 0 100 200 300 400 500
Coupling gap, s (nm) Coupling gap, s (nm)

Figure 1.15 Coupling strength kc as a function of the coupling gap


between two SOI waveguides for waveguide widths of 300, 400, and
500 nm: (a) TE polarization and (b) TM polarization. The waveguide
thickness is fixed at 250 nm. Solid lines are results obtained using semi-­
vectorial field distributions of the modes; dashed lines are results obtained
using full-vectorial field distributions.

M = Q Λ Q−1, where Λ is a diagonal matrix containing the eigenval-


ues of M and Q is a matrix containing its eigenvectors. The solution
to Equation 1.118 can then be expressed as

a( z) = Qe − jΛz Q −1a(0) = T( z)a(0), (1.124)

where a(0) represents the input fields [a1(0), a2(0)]T to the coupler at
z = 0, and T(z) is the transfer matrix of the coupler. The eigenvalues
of the matrix M in Equation 1.119 are found to be

λ = γ 0 ± Ω, (1.125)

where

Ω = δ 2 + k12 k 21 . (1.126)

In the above expressions, γ0 = (γ1 + γ2)/2 is the average propagation


constant and δ = (γ2 − γ1)/2 expresses the degree of asynchronism
or phase mismatch between the two waveguides. The eigenvector
matrix is

 k12 −(δ + Ω)


Q= . (1.127)
δ + Ω k21 
Elements of an Optical Microring Resonator 41

Substituting the eigenvalues and eigenvector matrix into Equation


1.124, we obtain the following expressions for the elements of the
transfer matrix T:
 δ 
T11 = T22 = cos(Ωz) + j sin(Ωz) e − jγ 0 z , (1.128)
 Ω 

k12
T12 = − j sin(Ωz)e − jγ 0 z , (1.129)

k 21
T21 = − j sin(Ωz)e − jγ 0 z . (1.130)

For an asymmetric coupler, δ2 > 0, Equation 1.126 shows that k12


and k21 are smaller than Ω so that |T21|2 < 1 and |T12|2 < 1. Thus
power from one waveguide can never be completely transferred
to the other. When the two waveguides are identical, δ = 0 and
Ω = k12 = k21 = kc; in this case, complete power transfer between the
two waveguides is achieved at the coupling length Lc = π/2kc.
For a symmetric coupler of length L, the transfer matrix is
given by
 cos(kc L) − j sin(kc L) − jγ 0 L
T= e . (1.131)
 − j sin( kc L) cos(kc L) 

In the analysis of coupled waveguide devices, it is often convenient


to model the coupler by a lumped coupling junction connected by
two uncoupled waveguides with length L and propagation constant
γ0, as shown in Figure 1.16. The transfer matrix of the lumped cou-
pling junction is given by
 τ − jκ 
T= , (1.132)
 − jκ τ 

Uncoupled
Distributed coupler waveguides Lumped coupler
τ γ0
β1 γ0
–jκ –jκ
β2 γ0 γ0
τ
L L/2 L/2

Figure 1.16 Representation of a pair of evanescently coupled waveguides


of length L by a lumped coupler connected by two uncoupled waveguides.
42 Optical Microring Resonators

where τ = cos(kcL) and κ = sin(kcL) are the field transmission coef-


ficient and field coupling coefficient, respectively, of the coupling
junction. Power conservation requires that τ2 + κ2 = 1. Note that in
an asymmetric coupler, the transmission coefficient has a small
phase shift as indicated by Equation 1.128.
When a straight waveguide is coupled to a curved waveguide,
the analysis is more complicated since the phase front tilt of the
bent waveguide mode must be taken into account. To simplify the
problem, we may approximate the curved waveguide by a series
of waveguide segments that are parallel to the straight waveguide,
as depicted in Figure 1.17b. This approximation is reasonable since
the coupling strength between two waveguides decreases exponen-
tially with the gap separation between them, so that the interac-
tion between a straight and bent waveguide quickly drops off as
the mismatch between their phase fronts becomes more significant.
The above simplification allows us to apply the result for the straight
waveguide coupler to compute the coupling coefficient between
a curved waveguide and a straight waveguide, or between two
curved waveguides. The approach, however, neglects the skewed
mode distribution of the bent waveguide and assumes that it can be
approximated by that of a straight waveguide with the same width.
Consider the coupling between a straight waveguide and a bent
waveguide with inner radius R1 and outer radius R2, as shown in
Figure 1.17. We denote the edge-to-edge gap separation between the
two structures by the function

s(θ) = s0 + R2 (1 − cos θ), (1.133)

(a) (b)
x x

R1 R2 R2
θ

s0 s0 s(θ)
dz
z z
0 0

Figure 1.17 (a) Evanescent coupling between a straight waveguide and a


bent waveguide. (b) Approximation of the bent waveguide by a series of
waveguide segments parallel to the straight waveguide.
Elements of an Optical Microring Resonator 43

with the minimum separation s0 occurring at θ = 0. Over each


s­ egment dz, the field coupling between the two parallel waveguide
segments is given by

dκ( z) = sin( kc (s)dz) ≈ kc (s)dz , (1.134)

where kc(s) is the coupling strength between two straight wave-


guides separated by a gap s(θ). From Figure 1.15, we find that kc
typically has an exponential dependence on s of the form

kc (s) = kc (0)e − s/σ . (1.135)

Since z = R2 sin θ and dz = R2 cos dθ, we integrate Equation 1.134


over θ from −π/2 to π/2 to get the total field coupling between the
straight and bent waveguides:
π /2

κ= ∫
− π /2
kc (s)R2 cos θdθ. (1.136)

The above expression can also be used to estimate the coupling


coefficient between two microring waveguides. The only change is
in the expression for the gap separation s(θ). For two microrings
with the same outer radii R2, s(θ) is given by

s(θ) = s0 + 2R2 (1 − cos θ), (1.137)

where s0 is the minimum gap separation between the two


microrings.
The plots in Figure 1.18a and b show the power coupling coeffi-
cient κ2 as a function of the minimum coupling gap s0 between an SOI
microring and a straight waveguide, and between two SOI micror-
ings of the same radius. All waveguides have core dimensions of
250 nm thickness and 300 nm width. The four sets of curves in the
plots correspond to microring radius R = 2, 5, 10, and 20 μm for both
TE (solid lines) and TM polarization (dashed lines). The plots show
that the coupling coefficient decreases rapidly with increasing cou-
pling gap. For the same gap, the coupling is also smaller for smaller
microring radius since the interaction length is shorter. Comparison
between Figure 1.18a and b also shows that coupling between two
microrings is weaker than coupling between a microring and a
straight waveguide because the separation distance s(θ) between the
two ring waveguides increases faster in the former case.
44 Optical Microring Resonators

(a) (b)
0.4 0.4

Power coupling coefficient, κ2


Power coupling coefficient, κ2
0.35 R = 20 µm 0.35 R = 20 µm
0.3 0.3
0.25 10 µm 0.25 10 µm
0.2 5 µm 0.2
0.15 0.15
2 µm 5 µm
0.1 0.1
0.05 0.05 2 µm
0 0
200 300 400 500 600 200 300 400 500 600
Coupling gap, s0 (nm) Coupling gap, s0 (nm)

Figure 1.18 Plots of the power coupling coefficient κ2 vs. the minimum
coupling gap s0 between (a) a microring of radius R and a straight wave-
guide, and (b) between two microrings of the same radius R. Solid lines
are for TE polarization; dashed lines are for TM polarization.

1.4 Fabrication of Microring Resonators


Ring resonators were first realized in the GeO2-doped silica mate-
rial (Kominato et al. 1992, Suzuki et al. 1992). Due to the relatively
low-index contrast of the material system (Δn < 1%), early ring
­resonators were fabricated with very large radii (R > 1 mm) in order
to reduce the bending loss. Much more compact microring resona-
tors were subsequently demonstrated using waveguides with high
lateral index contrast such as AlGaAs/GaAs (Rafizadeh et al. 1997,
Absil et al. 2000) and polysilicon (Little et al. 1998). However, since
the high-index contrast significantly reduces the lateral coupling
strength, very narrow coupling gaps (~200 nm) between the micror-
ing and the bus waveguides were required. Such small coupling
gaps were difficult to fabricate using conventional photolithogra-
phy, so electron beam lithography was typically required to achieve
good device performance (Rafizadeh et al. 1997, Absil et al. 2000).
One way to avoid the need for patterning very small ­coupling gaps
is to use a vertical coupling configuration which had been demon-
strated earlier. (Suzuki et al. 1992). In a vertically coupled microring
device, the microring and the bus waveguides were formed in two
different high-index layers separated by a thin, low-index coupling
layer. Since the layers were either epitaxially grown, as in the case
of III–V semiconductor devices (Absil et al. 2001, Grover et al. 2001),
or deposited using a chemical vapor deposition process, as in the
case of glass-based devices (Little et al. 1999, 2004), their thicknesses
could be accurately controlled, which allows for better control of
Elements of an Optical Microring Resonator 45

the coupling between the microring and bus waveguides. Another


advantage of the vertical coupling configuration is that since the
microring resonators and the bus waveguides reside on different
layers, the material layer for the microrings could be separately
designed for the application of interest, for example, for lasing,
photodetection, or electrooptic modulation applications. On the
other hand, the fabrication of vertically coupled microring devices
involves more processing steps than laterally coupled devices,
which adversely impacts device quality and fabrication yield. With
electron beam lithography and deep UV lithography (Bogaerts
et al. 2005) becoming more economically accessible, most micror-
ing devices nowadays are fabricated based on the lateral coupling
configuration with lateral feature sizes accurately defined down to
100 nm.
Figure 1.19 depicts a typical process flow for fabricating
integrated optics devices including microring resonators using
­
electron beam or UV lithography. The process starts with a sub-
strate consisting of a high-index core layer (e.g., Si) residing on a

(a) Resist layer (b)

Core layer Core layer


Under-cladding layer Under-cladding layer

Support substrate Support substrate

Waveguide core
(c) (d) Over-cladding layer

Under-cladding layer Under-cladding layer

Support substrate Support substrate

Figure 1.19 Overview of the processing steps for fabricating integrated


optics devices: (a) spin-coat a resist layer on a waveguide s­ubstrate,
(b) expose device pattern and develop resist, (c) transfer device ­pattern
onto core layer by dry-etching, and (d) remove resist and deposit
over-­cladding layer.
46 Optical Microring Resonators

low-index buffer layer (e.g., SiO2). These layers are either epitaxially
grown (e.g., for III–V semiconductors), defined by ion implantation
(e.g., for SOI material), or deposited by chemical vapor deposition
(for glass-based devices). To define the microring resonators and
other device structures on the chip, a photoresist or electron beam
resist layer is first spin-coated on the wafer. The device pattern on
a mask is then imprinted onto the resist layer by electron beam or
UV exposure. Next, the resist is developed and the device pattern
is transferred onto the waveguide core layer using a dry etching
process such as reactive ion etching (RIE) or inductively coupled
plasma (ICP) etching. The resist layer is then washed away and the
wafer is coated with an over-cladding layer. Many applications also
require additional processing steps to fabricate micro-heaters or
electrodes and contact pads on the chip.
Due to random variations in the fabrication process, the fabri-
cated devices typically have dimensions that are slightly deviated
from the designed values. As a result, the measured response of the
device will in general be different from the target response. One of
the most pronounced effects of fabrication-induced variations is
the detuning of the resonant frequencies of a microring from the
designed values. From Equation 1.70 we find that the change in
the resonant wavelength due to a change in the average radius of
a microring is given by Δλ = (λ/R)ΔR. As an example, for a 10-μm
radius microring resonator operating around the 1.55 μm wave-
length, a 1-nm change in the average radius can induce a shift in
the resonant wavelengths by as much as 0.15 nm, or almost 20 GHz.
Since it is not possible to control fabrication variations to within
1 nm with the current fabrication technology, a post-fabrication
method is typically required to correct for the resonance detun-
ing of the microring. A common approach is to use a micro-heater
fabricated directly above the microring to thermo-optically tune its
resonant frequencies. Several methods for permanently trimming a
microring resonance have also been developed. These include UV
trimming of the refractive index of the cladding material (Chu et al.
1999), or in the case of silicon waveguides, permanently altering the
index of the silicon core by oxidation (Chen et al. 2011, Shen et al.
2011), amorphization (Bachman et al. 2013) or by inducing strain
and stress in the material (Schrauwen et al. 2008).
Another practical issue that must be considered in the
design of a microring device is polarization dependence. Due to
Elements of an Optical Microring Resonator 47

birefringence, the resonant frequencies of TE and TM polarized


light in the microring generally do not coincide with each other. In
addition, the coupling coefficient between the microring and a bus
waveguide, or between two microrings, is also different for the TE
and TM polarizations, as evident from Figure 1.18. This polarization
sensitivity tends to be more pronounced for microrings made of
high-index contrast materials. In general, it is not possible to design
the microring waveguide dimensions to completely eliminate the
polarization dependence in both the resonant frequencies and the
coupling coefficients over a given bandwidth. For this reason, to
achieve polarization insensitive operation, a polarization diversity
scheme must be used, which consists of splitting the input light into
two orthogonal polarization components and feeding each compo-
nent into a separate microring circuit independently optimized for
that polarization. The outputs of the two circuits are then combined
to form the total response of the device.
A further issue that must be considered in practical micror-
ing device applications is the variation in the resonant frequencies
due to fluctuations in the ambient temperature. This is caused by
the fact that the refractive indices of the waveguide materials are
temperature dependent. For applications requiring the ­resonant
­frequency of the microring to be stabilized, the chip is typically
placed on a thermoelectric cooler and its temperature actively con-
trolled and monitored by a temperature sensor. It is possible to
­stabilize the device temperature to less than ±0.1°C using a feed-
back control loop.

1.5 Summary
This chapter provides a brief review of the basic concepts in inte-
grated optics that are essential for understanding the operation
of a microring device and the various physical parameters influ-
encing its behaviors. In particular, a familiarity with the theory
of optical waveguides, the characteristics of whispering gallery
modes in bent waveguides, and the CMT describing evanescent
wave coupling is necessary for analyzing and designing microring
resonator devices in subsequent chapters of the book. With a view
on the practical realization and implementation of these devices,
we also provided an overview of the standard lithographic pro-
cess for fabricating microring resonators and other integrated
48 Optical Microring Resonators

optics components, and discussed several important issues that


must be considered in the practical applications of these devices.
In Chapter 2, we will develop basic models for analyzing the spec-
tral responses and determining various performance characteris-
tics of a microring resonator. These models will also be used in
subsequent chapters for analyzing and designing more advanced
microring devices.

References
Absil, P. P., Hryniewicz, H. V., Little, B. E., Johnson, F. G., Ritter, K.
J., Ho, P.-T. 2001. Vertically coupled microring resonators using
polymer wafer bonding. IEEE Photonics Technol. Lett. 13(1):
49–51.
Absil, P. P., Hryniewicz, J. V., Little, B. E., Wilson, R. A., Joneckis,
L. G., Ho, P.-T. 2000. Compact microring notch filters. IEEE
Photonics Technol. Lett. 12(4): 398–400.
Bachman, D., Chen, Z., Fedosejevs, R., Tsui, Y. Y., Van, V. 2013.
Permanent fine tuning of silicon microring devices by femto-
second laser surface amorphization and ablation. Opt. Express
21(9): 11048–11056.
Berglund, W., Gopinath, A. 2000. WKB analysis of bend losses in
optical waveguides. J. Lightwave Technol. 18(8): 1161–1166.
Bogaerts, W., Baets, R., Dumon, P., Wiaux, V., Beckx, S., Taillaert,
D., Luyssaert, B., Van Campenhout, J., Bienstman, P., Van
Thourhout, D. 2005. Nanophotonic waveguides in silicon-
on-insulator fabricated with CMOS technology. J. Lightwave
Technol. 23(1): 401–412.
Chen, C. J., Zheng, J., Gu, T., McMillan, J. F., Yu, M., Lo, G. Q., Kwong,
D. L., Wong, C. W. 2011. Selective tuning of high-Q silicon pho-
tonic crystal nanocavities via laser-assisted local oxidation.
Opt. Express 19(13): 12480–12489.
Cheng, Y., Lin, W., Fujii, Y. 1990. Local field analysis of bent graded-
index planar waveguides. J. Lightwave Technol. 8(10): 1461–1469.
Chiang, K. S. 1993. Review of numerical and approximate methods
for the modal analysis of general optical dielectric waveguides.
Opt. Quantum Electron. 26: S113–S134.
Elements of an Optical Microring Resonator 49

Chiang, K. S. 1996. Analysis of the Effective Index Method for the


vector modes of rectangular-core dielectric waveguides. IEEE
Trans. Microwave Theory Tech. 44(5): 692–700.
Chin, M. K., Ho, S. T. 1998. Design and modeling of waveguide-
coupled single-mode microring resonators. J. Lightwave Technol.
16(8): 1433–1446.
Chu, S. T., Pan, W., Sato, S., Taneko, T., Little, B. E., Kokubun, Y. 1999.
Wavelength trimming of a microring resonator filter by means
of a UV sensitive polymer overlay. IEEE Photonics Technol. Lett.
11(6): 688–690.
Chuang, S.-L. 1987a. A coupled mode formulation by reciprocity
and a variational principle. J. Lightwave Technol. 5(1): 5–15.
Chuang, S.-L. 1987b. A coupled-mode theory for multiwaveguide
systems satisfying the reciprocity theorem and power conser-
vation. J. Lightwave Technol. 5(1): 174–183.
Coale, F. S. 1956. A traveling-wave directional filter. IRE Trans.
Microwave Theory. Tech. 4(4): 256–160.
Feng, N.-N., Zhou, G.-R., Xu, C., Huang, W.-P. 2002. Computation
of full-vector modes for bending waveguide using cylindrical
Perfectly Matched Layers. J. Lightwave Technol. 20(11): 1976–1980.
Frey, B. J., Leviton, D. B., Madison, T. J. 2006. Temperature-dependent
refractive index of silicon and germanium. In SPIE Astronomical
Telescopes + Instrumentation, International Society for Optics
and Photonics, Orlando, FL, article no. 62732J.
Grover, R., Absil, P. P., Van, V., Hryniewicz, J. V., Little, B. E., King, O.,
Calhoun, L. C., Johnson, F. G., Ho, P.-T. 2001. Vertically coupled
GaInAsP-InP microring resonators. Opt. Lett. 26(8): 506–508.
Haus, H. A., Huang, W. P., Kawakami, S., Whitaker, N. A. 1987.
Coupled mode theory of optical waveguides. J. Lightwave
Technol. 5: 16–23.
Heiblum, M., Harris, J. H. 1975. Analysis of curved optical wave-
guides by conformal transformation. IEEE J. Quantum Electron.
QE-11(2): 75–83.
Hiremath, K. R., Hammer, M., Stoffer, R., Prkna, L., Ctyroky, J. 2005.
Analytic approach to dielectric optical bent slab waveguides.
Opt. Quantum Electron. 37(1–3): 37–61.
50 Optical Microring Resonators

Jackson, J. D. 1999. Classical Electrodynamics, 3rd Ed. Hoboken,


NJ: John Wiley & Sons.
Kakihara, K., Kono, N., Saitoh, K., Koshiba, M. 2006. Full-vectorial
finite element method in a cylindrical coordinate system for
loss analysis of photonic wire bends. Opt. Express 14(23):
11128–11141.
Knox, R. M., Toulios, P. P. 1970. Integrated circuits for the millimeter
through optical frequency range. In Symposium on Submillimeter
Waves, New York, pp. 497–516.
Koshiba, M. 1992. Waveguide Theory by the Finite Element Method.
London: Kluwer Academic Publishers.
Kominato, T., Ohmori, Y., Takato, N., Okazaki, H., Yasu, M. 1992.
Ring resonators composed of GeO2-doped silica waveguides.
J. Lightwave Technol. 10(12): 1781–1788.
Little, B. E., Chu, S. T., Absil, P. P., Hryniewicz, J. V., Johnson, F. G.,
Seiferth, F., Gill, D., Van, V., King, O., Trakalo, M. 2004. Very
high-order microring resonator filters for WDM applications.
IEEE Photonics Technol. Lett. 16(10), 2263–2265.
Little, B. E., Chu, S. T., Pan, W., Ripin, D., Kaneko, T., Kokubun,
Y., Ippen, E. 1999. Vertically coupled glass microring resona-
tor channel dropping filters. IEEE Photonics Technol. Lett. 11(2):
215–217.
Little, B. E., Foresi, J. S., Steinmeyer, G., Thoen, E. R., Chu, S. T., Haus,
H. A., Ippen, E. P., Kimerling, L. C., Greene, W. 1998. Ultra-
compact Si-SiO2 microring resonator optical channel dropping
filters. IEEE Photonics Technol. Lett. 10(4): 549–551.
Lui, W. W., Xu, C.-L. Xu, Hirono, T., Yokoyama, K., Huang, W.-P.
1998. Full-vectorial wave propagation in semiconductor ­optical
bending waveguides and equivalent straight waveguide
approximations. J. Lightwave Technol. 16(5): 910–914.
Malitson, I. H. 1965. Interspecimen comparison of the refractive
index of fused silica. J. Opt. Soc. Am. 55:1205–1209.
Marcatili, E. A. 1969a. Dielectric rectangular waveguide and direc-
tional coupler for integrated optics. Bell Syst. Tech. J. 48:2071–2102.
Marcatili, E. A. 1969b. Bends in optical dielectric guides. Bell Syst.
Tech. J. 48(7): 2103–2132.
Elements of an Optical Microring Resonator 51

Marcuse, D. 1969. Mode conversion caused by surface imperfections


of a dielectric slab waveguide. Bell Syst. Tech. J. 48:3187–3215.
Marcuse, D. 1972. Light Transmission Optics. New York: Van Nostrand
Reinhold.
Okamoto, K. 2000. Fundamentals of Optical Waveguides. San Diego:
Academic Press.
Rafizadeh, D., Zhang, J. P., Hagness, S. C., Taflove, A., Stair, K. A.,
Ho, S. T. 1997. Waveguide-coupled AlGaAs/GaAs microcavity
ring and disk resonators with high finesse and 21.6-nm free
spectral range. Opt. Lett. 22(16): 1244–1246.
Rahman, B. M. A., Davies, J. B. 1984. Finite-element solution of inte-
grated optical waveguides. J. Lightwave Technol. LT-2(5): 682–688.
Rivera, M. 1995. A finite difference BPM analysis of bent dielectric
waveguides. J. Lightwave Technol. 13(2): 233–238.
Rowland, D. R., Love, J. D. 1993. Evanescent wave coupling of whis-
pering gallery modes of a dielectric cylinder. Inst. Elect. Eng.
Proc. J. 140:177–188.
Prabhu, A. M., Tsay, A., Han, Z., Van, V. 2010. Extreme miniatur-
ization of silicon add-drop microring filters for VLSI photonics
applications. IEEE Photonics J. 2(3): 436–444.
Schrauwen, J., Van Thourhout, D., Baets R. 2008. Trimming of ­silicon
ring resonator by electron beam induced compaction and
strain. Opt. Express 16(6): 3738–3743.
Shen, Y., Divliansky, I. B., Basov, D. N., Mookherjea, S. 2011. Electric-
field-driven nano-oxidation trimming of silicon microrings
and interferometers. Opt. Lett. 36(14): 2668–2670.
Steinlechner, J., Krüger, C., Lastzka, N., Steinlechner, S., Khalaidovski,
A., Schnabel, R. 2013. Optical absorption measurements on
crystalline silicon test masses at 1550 nm. Classical Quantum
Gravity 30(9): 095007.
Suzuki, S., Shuto, K., Hibino, Y. 1992. Integrated-optic ring resona-
tors with two stacked layers of silica waveguide on Si. IEEE
Photonics Technol. Lett. 4(11): 1256–1258.
Tien, P. K. 1971. Light waves in thin films and integrated optics.
Appl. Opt. 10: 2395–2419.
52 Optical Microring Resonators

Vlasov, Y. A., McNab, S. J. 2004. Losses in single-mode silicon-on-


insulator strip waveguides and bends. Opt. Express 12: 1622–1631.
Xiao, J., Ma, H., Bai, N., Liu X., Sun, X. 2009. Full-vectorial analysis of
bending waveguides using finite difference method based on
H-fields in cylindrical coordinate systems. Opt. Commun. 282:
2511–2515.
Xu, C. L., Huang, W. P., Stern, M. S., Chaudhuri, S. K. 1994. Full-
vectorial mode calculations by finite difference method. IEE
Proc.: Optoelectron. 141(5): 281–286.
Xu, Q., Fattal, D., Beausoleil, R. G. 2008. 1.5-μm-radius high-Q ­silicon
microring resonators. Opt. Express 16: 4309–4315.
Yariv, A. 1973. Coupled-mode theory for guided-wave optics. IEEE J.
Quantum Electron. 9: 919–933.
Chapter 2

Analytical Models of a
Microring Resonator

In Section 1.2.2, we showed that as light travels around a micror-


ing resonator, the periodic boundary condition requiring the field to
be replicated after every round-trip gives rise to discrete longitudi-
nal resonant modes. At a resonant frequency, constructive interfer-
ence of the field causes a large amount of energy to build up in the
microring. In practice, this energy must be supplied to the resonator
by an external source and, for useful applications, one must also be
able to extract the light out of the microring. In practical implemen-
tations of microring devices, the coupling of light into and out of the
resonator is accomplished by evanescent wave coupling between the
microring and one or two external bus ­waveguides, as depicted in
Figure 1.1. A microring coupled to a single bus waveguide is com-
monly referred to as the all-pass configuration, whereas a microring
coupled to two bus waveguides is called the add-drop configuration.
The spectral characteristics of a microring resonator are very
similar to those of a Fabry–Perot (FP) resonator, even though the
former is a traveling-wave resonator, while the latter is a standing-
wave resonator. Specifically, the all-pass microring behaves like an
FP resonator with one partially reflecting mirror, while the add-
drop microring behaves like one with both partially reflecting
mirrors. The distinction between a standing-wave resonator and a
traveling-wave resonator is that the field at any point in the former
structure is the result of the interference between two waves travel-
ing in opposite directions, whereas the field inside a traveling-wave
resonator is that of either the forward- or the backward-traveling
wave alone. As a result, the field in a standing-wave resonator has
a spatially-dependent amplitude distribution, whereas the field
amplitude in a traveling-wave resonator is nearly uniform. An
important advantage of microring resonators over FP ­resonators is
that they do not require mirrors, which are ­generally difficult to
realize in integrated optics. In addition, due to the unidirectional
propagation of light in the microring, the o ­ utput light signals of a

53
54 Optical Microring Resonators

microring device appear at physically isolated ports from the input


light, which is desirable for most integrated optics applications.
In this chapter, we develop the analytical models that will allow
us to determine the transmission responses and other spectral char-
acteristics of a microring resonator. In general, there are two alternate
formalisms that can be used to analyze the response of a micror-
ing device: the power coupling formalism and the energy coupling
formalism. The microring models based on these formalisms will
also be ­useful for analyzing and designing more complicated device
configurations as well as more advanced applications of microring
resonators in later chapters. We will begin in Section 2.1 by exam-
ining the resonance characteristics of a microring resonator in the
absence of coupling to an external waveguide. In Section 2.2 we will
derive the transfer functions of a microring resonator coupled to one
and two external waveguides using the power coupling formalism.
Alternative descriptions of the all-pass and add-drop microring
resonators based on the energy coupling formalism will be given
in Section 2.3. Finally, Section 2.4 will establish the relationship
between the power coupling and energy coupling formalisms and
discuss the validity of each model for analyzing microring devices.

2.1 Resonance Spectrum of an


Uncoupled Microring Resonator
Important resonant characteristics of a microring resonator can
be determined by considering the simple structure of an isolated
microring without any coupling to external waveguides. In Section
1.2.2, we found that the resonant condition for a monochromatic
light wave traveling around the microring in either the clockwise or
counterclockwise direction is βθ = m, where βθ is the angular propa-
gation constant and m is the longitudinal resonant mode number.
By writing βθ = nr(2π/λ)Reff, where nr is the effective index of the
microring waveguide at the effective radial distance* Reff from the
center, we can express the resonant condition as†
2πReff = mλ/nr . (2.1)

* The effective radius Reff is defined in Equation 1.57.


† The transverse resonant mode number appears implicitly in this resonant con-
dition through the effective index nr and the effective radius Reff of the trans-
verse mode. See Equation 1.70 for an explicit resonant condition containing both
the transverse and longitudinal mode numbers.
Analytical Models of a Microring Resonator 55

The above expression states that resonance occurs whenever the


round-trip path length of the microring is equal to an integer
­multiple of the guided wavelength. Approximating the effective
radius by the average radius R of the microring, we obtain from
Equation 2.1 the free space wavelength λm of the resonant mode m:

2πnr R
λm = . (2.2)
m

The resonance spectrum of a microring resonator thus consists of


a comb of resonance peaks occurring at discrete wavelengths λm.
The spacing between two consecutive resonances is called the free
spectral range (FSR) of the resonator. The FSR can be computed
by observing that it is equal to the wavelength change required
for the round-trip phase of the resonator to undergo a 2π shift.
From the expression for the round-trip phase of the microring,
ϕrt = nr(2π/λ)2πR, we take the derivative with respect to the wave-
length to get

dφrt  dnr nr  2π
= −  2πR. (2.3)
dλ  dλ λ  λ

Since the group index of the microring waveguide is given by


ng = nr − λdnr/dλ, we can also write Equation 2.3 as

dφrt 2π
= − ng 2 2πR. (2.4)
dλ λ

The FSR is defined as the wavelength change ΔλFSR required to


obtain a round-trip phase change Δϕrt = 2π:

2π λ2
∆λ FSR = =− . (2.5)
dφrt /dλ 2πng R

In terms of frequency, the FSR is given by

c c 1
∆f FSR = − ∆λ FSR = = , (2.6)
λ 2
2πng R Trt

where Trt = 2πngR/c is the round-trip group delay (GD), that is,
the time it takes a pulse to travel around the microring. Equation
2.6 shows that the FSR is inversely proportional to the microring
56 Optical Microring Resonators

radius. Note that since the group index of the microring wave-
guide also exhibits frequency dispersion, the FSR will also depend
on the frequency in general. This effect can be observed in Figure
1.10a and b, which show that the wavelength spacings between
successive resonant modes are not constant. The frequency dis-
persion of the FSR of a microring can have a significant impact in
many applications. For example, it places an upper limit on the
wavelength conversion bandwidth of the four-wave mixing pro-
cess in a nonlinear microring resonator. This will be discussed in
detail in Chapter 4.
In Section 1.2.2, we also found that the field amplitude of a
resonant mode is constant along the direction of propagation (the
angular direction), implying that the power distribution is uniform
around the microring. From Equation 1.68 we may express the
amplitude of the resonant mode m as

A(t) = A0 e jω mt e − γ mt , (2.7)

where ωm is the resonant frequency and γm is the decay rate of the


amplitude of mode m. The amplitude decay rate is related to the
cavity lifetime* by γm = 1/2τm, or to the power attenuation constant α
of the microring waveguide through Equation 1.72,

ω m α/2 cα
γm = = . (2.8)
m/Reff 2nr

The amplitude A0 in Equation 2.7 is usually normalized with


respect to either energy or power. In the energy normalization, A0
is normalized such that |A0|2 gives the initial stored energy in the
microring and |A(t)|2 represents the stored energy at time t. In the
power normalization, |A0|2 and |A(t)|2 give the power flow through
any cross section of the microring waveguide at the initial time and
at time t, respectively.
To obtain the frequency response of the mode amplitude, we
take the Fourier transform of Equation 2.7 to get

A0
A(ω) = . (2.9)
j(ω − ω m ) + γ m

* Recall that we define the cavity lifetime (or photon lifetime) τm as the time it
takes the energy of mode m to decay to 1/e of its initial value.
Analytical Models of a Microring Resonator 57

The energy or power spectrum is obtained by taking the absolute


square of the above expression,
2
2 A0
A(ω) = . (2.10)
(ω − ω m )2 + γ m2

The above equation shows that the energy or power spectrum of the
microring has the form of a Lorentzian resonance centered around
the resonant frequency ωm. The full width at half max (FWHM)
bandwidth, or linewidth of the resonance, is obtained by equating
the expression in Equation 2.10 to half its peak value. The result is

ωm
∆ω BW = 2γ m = , (2.11)
Q0

where Q 0 = ωmτm is the intrinsic Q factor of mode m given in


Equation 1.77. The above relation shows that the resonance
linewidth is inversely proportional to the Q factor (or to the
cavity lifetime). It also gives the well-known formula for calcu-
lating the Q factor as the ratio of the resonant frequency to the
FWHM ­bandwidth, Q 0 = ωm/ΔωBW = λm/ΔλBW. Another useful
formula relating the intrinsic Q factor to the power attenuation
coefficient α of the microring waveguide is obtained by using
Equation 2.8 for γm:

2πnr
Q0 = . (2.12)
αλ m

2.2 Power Coupling Description


of a Microring Resonator
In this section, we determine the spectral response of a microring
resonator in the presence of coupling to one or two external bus
waveguides. There are two different approaches which we can use
to analyze such a structure: the power coupling approach (or “cou-
pling of modes in space”) and the energy coupling approach (or
“coupling of modes in time”). In the power coupling approach, the
device response is described in terms of power exchange between
the microring and the external bus waveguides. Couplings between
the microring and the bus waveguides are described by the field
coupling coefficients, which are obtained from the solution of the
58 Optical Microring Resonators

coupled mode equations in space (Section 1.3.2). In the energy cou-


pling approach, the behavior of the device is described from the
point of view of energy exchange between the microring and the
external bus waveguides. Strictly speaking, the energy coupling
formalism is valid only for weakly coupled microring resonators,
whereas there is no restriction on the coupling strengths in the
power coupling approach. In this section we will use the power
coupling approach to derive the transfer functions of a microring
resonator in the add-drop and all-pass configurations. Analysis of
these structures based on the energy coupling formalism will be
given in Section 2.3.

2.2.1 The add-drop microring resonator


An add-drop microring resonator (ADMR) is a four-port optical
device which consists of a microring evanescently coupled to an
input waveguide and an output waveguide, as shown in Figure 2.1a.
The four ports are labeled input, through, drop, and add. An opti-
cal signal applied to the input port with frequency coinciding to a
microring resonance will be transmitted (or “dropped”) at the drop
port. If the frequency of the input signal is tuned off resonance, the
light will instead be transmitted to the through port. The device can
thus be used as an add-drop filter in a wavelength division multi­
plexing (WDM) communication network. It allows a wavelength
channel from an input WDM signal that is in resonance with the
microring to be extracted to the drop port while allowing all other

(a)
si st
Input port Through port
κ1, τ1
Input waveguide (b)
D A τ1 τ2
R
si sd
C B
Output waveguide st
κ2, τ2
Drop port Add port
sd sa

Figure 2.1 (a) Schematic of an ADMR with average radius R, input cou-
pling coefficient κ1 and output coupling coefficient κ2. (b) In terms of
transfer characteristics, the ADMR is equivalent to an FP resonator with
mirror reflectivities τ1 and τ2.
Analytical Models of a Microring Resonator 59

channels to bypass to the through port. Conversely, a channel that


is in resonance with the microring can be added onto the incoming
WDM stream via the add port.
To derive the transfer functions at the drop port and through
port of an ADMR, we consider the microring in Figure 2.1a with
average radius R evanescently coupled to an input and an out-
put waveguide. We represent each coupling junction between the
microring and a bus waveguide as a lumped coupler with field cou-
pling coefficient κi and transmission coefficient τi, where i = 1 for
the input and i = 2 for the output. For simplicity we assume the cou-
pling junctions to be lossless, so that κ i2 + τ i2 = 1.* Suppose a mono-
chromatic light wave of frequency ω is applied to the input port. Its
electric field can be expressed as

Ei (r , t) = E0 e( x , y )e j(ωt −β b z ) , (2.13)

where e(x, y) is the modal field distribution normalized to unit


power, βb is the propagation constant of the bus waveguide, and
z is the direction of propagation along the input waveguide. Let si
represent the complex amplitude of the input signal at a point z0
just before the input coupling junction, si = E0 e − jβ b z0 . Note that si is
normalized such that the power in the input waveguide is given by
|si|2 = |E0|2. In a similar manner we denote sa as the amplitude of
the light wave applied to the add port, and st and sd as the output
signals at the through port and drop port, respectively.
At the input and output coupling junctions, the applied signals
at the input and add ports are coupled into the microring, giving
rise to a clockwise-propagating wave in the resonator. We label the
complex amplitude of this wave at the points just before and after
the input and output coupling junctions as A, B, C, and D, as shown
in Figure 2.1a. Using the transfer matrix in Equation 1.132 for the
input coupler, we have

st = − jκ 1D + τ1si , (2.14)

A = − jκ 1si + τ1D. (2.15)

* To account for coupling junction loss, we replace the coupling and transmission
coefficients by α i κ i and α i τ i, respectively, such that α iκ i2 + α i τ i2 = α i, where αi
is the power attenuation due to the junction loss of coupler i.
60 Optical Microring Resonators

Assuming that there is no applied signal at the add port, sa = 0,


we also obtain the following relations at the output coupling
junction:

sd = − jκ 2 B, (2.16)

C = τ 2 B. (2.17)

The fields B and D can be related to the fields A and C, respectively,


by

B = Ae − jγπR = A art e − jφrt /2 , (2.18)

D = Ce − jγπR = C art e − jφrt /2 , (2.19)

where γ = β − jα/2 is the complex propagation constant of the micror-


ing waveguide, ϕrt = β2πR is the round-trip phase, and art = e −απR is
the round-trip field attenuation factor in the micro­ring. The propa-
gation loss constant α accounts for the total waveguide loss, includ-
ing absorption, bending loss, and scattering. Using Equations 2.18
and 2.19 to eliminate C and B in Equation 2.17, we get

D = τ 2 Aart e − jφrt . (2.20)

Substituting the above result into Equation 2.15 and solving for A,
we obtain

− jκ 1si
A= . (2.21)
1 − τ1τ 2 art e − jφrt

Similar expressions for B, C, and D can also be derived in terms of si


using Equations 2.17 through 2.19. These signals are ­approximately
equal to A in magnitude if the microring loss and coupling coef-
ficients are small (art ≈ 1, τ1 ≈ 1, τ2 ≈ 1). For these high-Q ­resonators,
the field amplitude can be assumed to be uniformly distributed
around the microring.* However, for strongly coupled or high-loss
microring resonators, the field amplitude distribution will gener-
ally be nonuniform.

* This is in contrast to a FP resonator, where the field amplitude distribution is


determined by the standing-wave pattern in the cavity.
Analytical Models of a Microring Resonator 61

We note from Equation 2.21 that the field amplitude |A| is max-
imum at the resonant wavelengths λm, where the round-trip phase
ϕrt = 2mπ. We can define the field enhancement factor, FE, as the
ratio of the peak field amplitude inside the microring at resonance
to the input field amplitude:

A κ1
FE = = . (2.22)
si φrt = 2 mπ 1 − τ1τ 2 art

If the microring is symmetrically coupled (τ1 = τ2) and has low loss
(art ≈ 1), the above equation gives FE ≈ 1/κ, that is, the field enhance-
ment factor is inversely proportional to the field coupling coeffi-
cient. Thus, by choosing a small coupling coefficient (κ ~ 0.1–0.01),
we can amplify the field circulating in the microring by 1–2 orders
of magnitude with respect to the input light wave. This resonance
field enhancement has been widely exploited to reduce the power
requirement in nonlinear optics applications of microring resona-
tors, such as all-optical switching and frequency conversion. We
will discuss some of these applications in Chapter 4.
To determine the transfer function at the drop port of the
ADMR, we substitute B = A art e − jφrt /2 into Equation 2.16 and, using
Equation 2.21, obtain the following result:

sd κ κ a e − jφrt /2
Hd (φrt ) ≡ = − 1 2 rt − jφrt . (2.23)
si 1 − τ1τ 2 art e

Similarly, substituting D = τ 2 Aart e − jφrt into Equation 2.14 and using


Equation 2.21, we obtain the transfer function at the through port
as

st τ1 − τ 2 art e − jφrt
H t (φrt ) ≡ = . (2.24)
si 1 − τ1τ 2 art e − jφrt

In arriving at the above result, we have also made use of the relation
τ12 + κ 12 = 1. If a signal sa is applied to the add port of the ADMR and
there is no input signal (si = 0), the transfer functions at the through
port (st/sa) and drop port (sd/sa) are also given by Equations 2.23 and
2.24, respectively, with the coefficients τ1 (κ1) and τ2 (κ2) interchanged.
The ADMR can be regarded as an infinite impulse response
(IIR) digital filter (Madsen and Zhao 1999, Chapter 3) since the out-
put signal at the drop port of the microring can be expressed as an
62 Optical Microring Resonators

infinite series of signals, with each successive signal delayed by one


round-trip time. Indeed, by defining the round-trip delay variable
z −1 = e − jφrt, we can express the transfer functions in Equations 2.23
and 2.24 as

κ 1κ 2 art z −1/2
Hd ( z −1 ) = − , (2.25)
1 − τ1τ 2 art z −1

τ1 − τ 2 art z −1
H t ( z −1 ) = . (2.26)
1 − τ1τ 2 art z −1

The above expressions have the form of a transfer function of a digi-


tal filter with unit delay variable z−1. Note that both are first-order
transfer functions with a pole at z−1 = 1/τ1τ2 art. Additionally, the
through port transfer function has a zero at z−1 = τ1/τ2 art, while the
drop port transfer function has a trivial zero at z−1 = 0. By taking
the inverse z-transforms of Hd and Ht, we can obtain the discrete-
time impulse responses of the ADMR at the drop port and through
port, respectively.
The power spectral responses at the drop port and through
port of the ADMR can be obtained by taking the absolute square of
Equations 2.23 and 2.24:
2
sd κ 12κ 22 art
Td = = , (2.27)
si 1 + τ12 τ 22 art2 − 2τ1τ 2 art cos φrt

2
st τ12 + τ 22 art2 − 2τ1τ 2 art cos φrt
Tt = = . (2.28)
si 1 + τ12 τ 22 art2 − 2τ1τ 2 art coss φrt

Figure 2.2 shows representative plots of the spectral responses at


the drop port and through port of an ADMR as functions of the
round-trip phase ϕrt. We see that the drop port response has trans-
mission peaks at the resonances, while the through port response
exhibits corresponding transmission dips. By substituting ϕrt = 2mπ
into Equations 2.27 and 2.28, we obtain the values for the maximum
transmission at the drop port, Td,max, and the minimum transmis-
sion at the through port, Tt,min, as follows:

κ 12κ 22 art κ 12κ 22 art


Td ,max = = , (2.29)
1 + τ12 τ 22 art2 − 2τ1τ 2 art (1 − τ1τ 2 art )2
Analytical Models of a Microring Resonator 63

0
∆φBW Insertion

Transmission (dB)
–5 loss

–10 Td

–15 Tt

–20 FSR Extinction


–π 0 π 2π 3π
Roundtrip phase detune (∆φrt)

Figure 2.2 Power spectral responses at the drop port (black line) and
through port (gray line) of an ADMR with power coupling coefficients
κ 12 = κ 22 = 0.2 and 5% round-trip power loss (art = 0.975).

2
τ12 + τ 22 art2 − 2τ1τ 2 art  τ1 − τ 2 art  (2.30)
Tt ,min = = .
1 + τ12 τ 22 art2 − 2τ1τ 2 art  1 − τ1τ 2 art 

Using the above expressions, we can express the power spectral


responses in Equations 2.27 and 2.28 as

Td ,max
Td (φrt ) = , (2.31)
1 + F sin 2 (φrt /2)

Tt ,min + F sin 2 (φrt /2)


Tt (φrt ) = , (2.32)
1 + F sin 2 (φrt /2)

4τ1τ 2 art
F= . (2.33)
(1 − τ1τ 2 art )2

The parameter F is called the contrast of the resonator and provides


a measure of the sharpness of the resonance peak. Equation 2.31
shows that the drop port response of an ADMR has the form of an
Airy function, which is characteristic of a FP resonance. Indeed,
the drop port and through port transfer functions of an ADMR are
formally identical to those of an FP resonator with mirror reflec-
tivities τ1 and τ2, as shown in Figure 2.1b. Note, however, that in
the ADMR the transmitted signal at the through port is physically
isolated from the input signal instead of being reflected back into
the input port, as in the case of the FP resonator. The fact that there
64 Optical Microring Resonators

is no back-reflected signal* at the input port of an ADMR makes it


more attractive than FP resonators for photonic integrated circuit
applications.
For a lossless and symmetrically coupled microring resonator
(art = 1, τ1 = τ2), Equations 2.29 and 2.30 give Td,max = 1 and Tt,min = 0.
Thus, at the resonant frequencies, all the input power is transmitted
to the drop port and complete extinction is achieved at the through
port. In this case the field buildup in the microring is such that
the fraction of light coupled out onto the input waveguide inter-
feres destructively with the input signal, resulting in zero trans-
mitted power at the through port. Since power must be conserved
in the absence of loss, the input power is completely transferred
to the drop port. This process is called resonance-assisted power
transfer between the input and output waveguides. Note that com-
plete power transfer is possible even if there is phase mismatch
between the bus waveguides and the microring waveguide.† The
only requirement for complete power transfer in a lossless ADMR
is the matching of the input and output coupling coefficients.
When loss is present in the microring, the peak drop port trans-
mission is always less than unity. The difference in decibels between
the peak power transmission and unity is called the insertion loss
of the device. Complete extinction at the through port can still be
achieved in the presence of loss by a suitable choice of the coupling
coefficients. In particular, we find from Equation 2.30 that Tt,min = 0
if τ1 = τ2 art. This condition, which results in zero transmission at the
through port, is called critical coupling in an ADMR.
To determine the FWHM (or 3 dB) bandwidth ΔϕBW of the
ADMR, we equate the drop port power transmission in Equation
2.31 to half its peak value,

Td ,max T
Td (∆φ BW /2) = = d ,max , (2.34)
1 + F sin 2 (∆φ BW /4) 2

which gives

* In practice, there is a small back-reflected signal at the input port of an ADMR


due to backscattering from waveguide imperfections and the coupling junc-
tions; however, this reflected signal can be minimized with proper design and
an optimized fabrication process.
† Recall that in an evanescent wave coupler, complete power transfer is not pos-
sible if there is phase mismatch between the two waveguides, that is, if they
have different propagation constants.
Analytical Models of a Microring Resonator 65

sin( ∆φ BW /4) = 1/ F . (2.35)

Note that the above expression can also be obtained from the
through port power transmission in Equation 2.32 by equating
Tt(ΔϕBW/2) = (1 + Tt,min)/2. For devices with narrow bandwidths,
we can use the small-angle approximation of the sine function in
Equation 2.35 to get

4 2(1 − τ1τ 2 art )


∆φ BW ≈ = . (2.36)
F τ1τ 2 art

Since ΔϕBW = ΔωBWTrt,* where Trt is the round-trip GD, we also have
the following expressions for the FWHM bandwidths in terms of
frequency and wavelength:

1 ∆φ BW vg vg 1 − τ1τ 2 art
∆f BW = = 2 = 2 , (2.37)
2π Trt π R F 2π R τ1τ 2 art

λ 2m λ 2m 1 − τ1τ 2 art
∆λ BW = = 2
. (2.38)
π 2 ng R F 2π ng R τ1τ 2 art

In the above equations, vg and ng are the group velocity and group
index, respectively, of the microring waveguide.
Using the expression for the bandwidth ΔλBW in Equation 2.38,
we can determine the loaded quality factor of the ADMR as†

λm π 2 ng R F 2π 2 ng R τ1τ 2 art
Q= = = . (2.39)
∆λ BW λm λ m 1 − τ1τ 2 art

For low-loss microrings (art ≈ 1) the loaded Q factor is dominated by


external couplings to the bus waveguides. If the input and output

* Since the round-trip phase of the microring is ϕrt = nr(ω/c)2πR, we have


dφrt  nr dnr ω   dn  2πR 2πR
= + 2πR =  nr + ω r  = = Trt .
dω  c dω c   dω  c c / ng
Using the above result, we get ΔϕBW = ΔωBWTrt.
† Strictly speaking, the definition of Q as λm divided by the FWHM bandwidth is
valid only for classical oscillators of the Lorentzian type (see Equation 2.11). The
expression in Equation 2.39 provides a good estimate of the Q factor for low loss
and weakly coupled microring resonators (art ≈ 1 and τ1 ≈ τ2 ≈ 1).
66 Optical Microring Resonators

coupling coefficients are assumed to be equal (τ1 = τ2 = τ) and small


(τ ≈ 1), we can approximate the Q factor as

2π 2 ng R τ 2π 2 ng R
Q≈ ≈ . (2.40)
λm 1 − τ2 λ mκ 2

The above expression shows that the loaded Q factor increases with
the microring radius R and is inversely proportional to the power
coupling coefficient κ2 and the resonant wavelength λm.
We can also determine the finesse of the ADMR by dividing
the FSR of the microring in Equation 2.5 by the 3 dB or FWHM
bandwidth,

∆λ FSR π π τ1τ 2 art


ℱ= = F= . (2.41)
∆λ BW 2 1 − τ1τ 2 art

For a symmetrically coupled ADMR with low loss, the above


expression simplifies to ℱ ≈ π/κ 2. Since FE ≈ 1/κ, we find that the
finesse is related to the field enhancement in the microring by
F = π(FE)2. From Equations 2.39 and 2.41 we also obtain the relation-
ship between the finesse and the Q factor,
λ mQ Q
ℱ= ≈ , (2.42)
2πng R m

where we have used the expression in Equation 2.2 for the resonant
wavelength λm and assumed that nr ≈ ng.
The physical meaning of the finesse can be deduced by noting
that the cavity lifetime of the microring is inversely proportional to
the bandwidth ΔωBW (e.g., see Equation. 2.11):

1 T F Trt ℱ
τc = = rt = . (2.43)
∆ω BW 4 2π

The above result shows that the finesse (divided by 2π) gives the
number of round-trips a light wave makes around the microring
before the stored energy decays to 1/e of the initial value due to loss
and external coupling. In other words, the finesse gives the cavity
lifetime in terms of the number of round-trips. Since there are m
periods of wave oscillations per round-trip for resonant mode m,
we see from Equation 2.42 that the Q factor gives the cavity lifetime
in terms of the number of periods of oscillations.
Analytical Models of a Microring Resonator 67

2.2.2 The all-pass microring resonator


An all-pass microring resonator (APMR) consists of a microring
coupled to a single bus waveguide with field coupling coefficient
κ, as shown in Figure 2.3a. The structure can be regarded as a spe-
cial case of the add-drop configuration with τ2 = 1. Thus, by setting
τ1 = τ and τ2 = 1 in Equation 2.24, we obtain the transfer function of
the APMR at the output (or through) port as follows:

st τ − art e − jφrt τ − art z −1


H ap ≡ = − jφrt
= . (2.44)
si 1 − τart e 1 − τart z −1

The power spectral response is given by


2
τ 2 + art2 − 2τart cos φrt Tap, min + F sin (φrt /2)
Tap (φrt ) = = , (2.45)
1 + τ 2 art2 − 2τart cos φrt 1 + F sin 2 (φrt /2)

4τart
F= , (2.46)
(1 − τart )2

2
 τ − art  (2.47)
Tap,min =  .
 1 − τart 

The spectral characteristics of an APMR are similar to those of


an FP resonator with a perfectly reflecting mirror and a partially
reflecting mirror with reflectivity τ, as shown in Figure 2.3b. In the
absence of loss (art = 1), Tap,min = 1 and Equation 2.45 gives Tap = 1,

(a) (b)
τ τ=1
R
si Perfectly
B A reflecting
st mirror
κ, τ
Input port Output port
si st

Figure 2.3 (a) Schematic of an APMR with field coupling coefficient κ.


(b) In terms of transfer characteristics, the APMR is equivalent to an FP
resonator with a perfectly reflecting mirror and a partially reflecting
­mirror with reflectivity τ.
68 Optical Microring Resonators

indicating that power is transmitted through the device with-


out attenuation at all frequencies. The microring thus behaves as
an all-pass filter. In the presence of loss, the spectral response is
characterized by a transmission dip at the resonant wavelengths.
In particular, when τ = art, Equation 2.47 gives Tap,min = 0, implying
that total extinction of the output signal is achieved at the resonant
wavelengths. The microring resonator is said to be critically cou-
pled (Yariv 2000). Physically, the input light is destructively inter-
fered by the light coupled out from the microring, resulting in zero
power being transmitted at the output port. All the input power
in this case is dissipated through the various loss mechanisms in
the microring. For other values of τ and art, the input power is only
partially dissipated and the remaining power is transmitted to the
output port. For this more general case, the microring is said to
be undercoupled if τ > art and overcoupled if τ < art. Typical power
spectral responses of an APMR in the regimes of undercoupling,
overcoupling, and critical coupling are shown in Figure 2.4a. The
associated power enhancements in the microring, |FE|2 = |A/si|2,
where A is the field in the microring given by Equation 2.21, are
also plotted in Figure 2.4b. We observe that among the different
coupling regimes, the critically-coupled microring has the deep-
est extinction and the largest power enhancement at resonance.
We also note that overcoupled APMRs have broader bandwidths
than undercoupled microrings because of stronger couplings to the
bus waveguide.

(a) (b)
1 12
Critically coupled
Power enhancement (|FE|2)

0.8 10 τ = 0.95 Overcoupled


Transmission (Tap)

τ = 0.9 τ = 0.9
τ = 0.8 8 τ = 0.8
0.6 τ = 0.7 τ = 0.7
Undercoupled Overcoupled 6 Undercoupled
0.4 τ = 0.99 τ = 0.97
τ = 0.98 4 τ = 0.98
τ = 0.97 Critically τ = 0.99
0.2 coupled 2
τ = 0.95
0 0
–0.2π –0.1π 0 0.1π 0.2π –0.2π –0.1π 0 0.1π 0.2π
Roundtrip phase detune (∆φrt) Roundtrip phase detune (∆φrt)

Figure 2.4 (a) Power spectral responses of an APMR for various coupling
values and fixed round-trip attenuation art = 0.95. (b) Power enhancement
|FE|2 in the microring as a function of the round-trip phase detune.
Analytical Models of a Microring Resonator 69

The FWHM bandwidth* of the resonance spectrum of an


APMR can be obtained by setting τ1 = τ and τ2 = 1 in Equation 2.38,

λ 2m 1 − τart
∆λ BW = . (2.48)
2π 2 ng R τart

For low-loss (art ≈ 1) and weakly coupled (τ ≈ 1) microrings, we can


approximate the above expression as follows:

λ 2m 1 − τ λ 2m 1 − τ2 λ m2 κ 2
∆λ BW ≈ ≈ ≈ . (2.49)
2π 2 ng R τ 2π 2 ng R τ (1 + τ) 4π 2 ng R

The loaded Q factor and the finesse of the APMR are given by

2π 2 ng R τart 4π 2 ng R
Q= ≈ , (2.50)
λ m 1 − τart λ mκ 2

π τart 2π
ℱ= ≈ , (2.51)
1 − τart κ 2

where the approximations again apply to low-loss and weakly cou-


pled microrings. Comparing Equations 2.40 and 2.50, we find that
the Q factor (and also the finesse) of an APMR is twice that of an
ADMR with the same coupling coefficient. This result is expected
since an APMR has half the loss due to external coupling as an ADMR.

2.2.3 Phase and GD responses of a microring resonator


The phase response and GD response of an APMR or ADMR are
determined by the poles and zeros of the transfer functions of the
respective device. Consider first the transfer function at the drop
port of an ADMR (Equation 2.23), which can be written as

κ 1κ 2 art e − jφrt /2
Hd (φrt ) = − , (2.52)
1 − p0−1e − jφrt

* For a resonance dip, the FWHM bandwidth is the bandwidth measured halfway
between the maximum transmission (off resonance) and the transmission dip
(at the resonant frequency).
70 Optical Microring Resonators

where p0 = 1/τ1τ2 art is the pole in the inverse z-plane ( z −1 = e − jφrt ).


The phase response of Hd is given by

φrt  sin φrt 


ψ d (φrt ) = π − − tan −1  , (2.53)
2  p0 − cos φrt 

where the last term is the phase response due to the pole. The GD
response is obtained by differentiating the phase ψ of the transfer
function with respect to the angular frequency, τg = −dψ/dω. Since
dϕrt/dω = Trt, we have

dψ dφ dψ dψ
τg = − = − rt = −Trt . (2.54)
dω dω dφrt dφrt

The above expression shows that a pulse passing through the


microring device experiences a GD that is enhanced by a factor
S = −dψ/dϕrt over the round-trip delay Trt of the microring wave-
guide. The parameter S is called the GD enhancement factor or the
slowness factor. By differentiating the phase ψd in Equation 2.53
with respect to ϕrt and using the result in Equation 2.54, we obtain
the GD response at the drop port of the ADMR as follows:

( p02 − 1)Trt /2
τ g ,d (φrt ) = . (2.55)
1 + p02 − 2 p0 cos φrt

Writing cos ϕrt = 1 − 2 sin2(ϕrt/2), we can also express the above


equation in the form

Sp, max
τ g,d (φrt ) = Trt , (2.56)
1 + F sin 2 (φrt /2)

where F = 4p0/(p0 − 1)2 = 4τ1τ2 art/(1 − τ1τ2 art)2 is the contrast param-
eter defined in Equation 2.33, and

1  p0 + 1  1  1 + τ1τ 2 art 
Sp, max = = (2.57)
2  p0 − 1  2  1 − τ1τ 2 art 

is the maximum GD enhancement factor. Equation 2.56 shows that


the GD response has the same shape as the power spectral response
at the drop port (Equation 2.31). Note that the GD at the drop port
Analytical Models of a Microring Resonator 71

is always positive. Physically, the parameter Sp,max gives the number


of round-trips a pulse makes around the microring at resonance
before exiting at the drop port. Like the finesse, it provides a m
­ easure
of the cavity lifetime in terms of the number of round-trips.*
For the through port of the ADMR, we write the transfer
­function in Equation 2.24 as

1 − (τ 2 art /τ1 )e − jφrt 1 − z0−1e − jφrt


H t (φrt ) = τ1 = τ 1 , (2.58)
1 − τ1τ 2 art e − jφrt 1 − p0−1e − jφrt

where z0 = τ1/τ2art is the zero and p0 = 1/τ1τ2art is the same pole appear-
ing in the drop port transfer function. The phase response at the
through port consists of contributions from both the zero and the pole,

 sin φrt   sin φrt 


ψ t (φrt ) = tan −1   − tan −1  . (2.59)
 z0 − cos φrt   p0 − cos φrt 

To obtain the GD response, we write Equation 2.59 as

φ  sin φrt    φrt −1  sin φrt  


ψ t (φrt ) =  rt + tan −1   −  + tan  p − cos φ   .
 2  z0 − cos φrt    2 0 rt 

(2.60)

Noting the resemblance of the terms in the square brackets to the


phase response in Equation 2.53, we can write the GD response at
the through port as

( z02 − 1)Trt /2 ( p02 − 1)Trt /2


τ g,t (φrt ) = − + . (2.61)
1 + z02 − 2z0 cos φrt 1 + p02 − 2pp0 cos φrt

The first term on the right-hand side gives the delay due to the zero,
while the second term is the delay due to the pole. Equation 2.61 can
also be written as

Sz ,max Sp ,max
τ g,t (φrt ) = − 2
Trt + Trt = − τ g,zero + τ g,pole ,
1 + G sin (φrt /2) 1 + F sin 2 (φrt /2)
(2.62)

* For high-Q microring resonators (art ≈ 1, τ1 ≈ τ2 ≈ 1), Equation 2.57 gives


Sp,max ≈ (1 + τ2)/2κ2 ≈ 1/κ2. Since ℱ ≈ π/κ2, we have Sp,max ≈ ℱ/π.
72 Optical Microring Resonators

1  z + 1  1  τ1 + τ 2 art 
Sz, max =  0 = , (2.63)
2  z0 − 1  2  τ1 − τ 2 art 

4 z0 F
G= 2
= , (2.64)
( z0 − 1) Tt, min

where Tt,min is the minimum power transmission at the through


port given by Equation 2.30. Equation 2.62 shows that the effect of
the pole is to add a positive delay, whereas the effect of the zero is to
introduce a negative delay if τ1 > τ2 art. At the resonant frequencies,
the total GD is positive if τ1 < τ2 art (overcoupling) and negative if
τ1 > τ2 art (undercoupling). When τ1 = τ2 art (critical coupling), the GD
due to the zero, τg,zero, approaches negative infinity at the resonant
­frequencies. However, the power transmitted at the through port
also becomes vanishingly small in this case.
For an APMR, the phase and GD responses are given by the same
expressions as those at the through port of an ADMR (Equations
2.59 and 2.61), with the pole and zero replaced by p0 = 1/τart and
z0 = τ/art, respectively. Similar to the through port response of an
ADMR, the GD of an APMR is infinite at the resonant frequencies
for the critical coupling case (τ = art), positive for the undercoupling
case (τ > art), and negative for the overcoupling case (τ < art).
Figure 2.5a and b show the phase and GD responses, respec-
tively, at the drop port and through port of a symmetrically coupled
ADMR (τ1 = τ2). In the presence of loss, the device is undercoupled
since τ1 > τ2 art. We observe in Figure 2.5a that the phases at both

(a) 1.5π (b)


10
Drop port
5
GD enhancement (S)

π 0
Phase response (ψ)

–5
Drop port
0.5π –10
Through port
–15
Through port
0 –20
–25
–0.5π –30
–π 0 π 2π 3π –0.5π –0.25π 0 0.25π 0.5π
Roundtrip phase detune (∆φrt) Roundtrip phase detune (∆φrt)

Figure 2.5 (a) Phase response and (b) group delay response (S = τg/Trt)
at the drop port and through port of an ADMR with coupling coefficients
κ 12 = κ 22 = 0.2 and 5% round-trip power loss (art = 0.975).
Analytical Models of a Microring Resonator 73

the drop port and through port undergo sharp transitions at the
resonances. This rapid phase change near a resonant frequency can
be exploited to enhance the performance of phase-sensitive or inter-
ferometric devices such as switches, modulators, and sensors. Note
that the slope of the phase response at resonance is negative at the
drop port and positive at the through port. As a result, we obtain a
positive-peak GD response at the former and a negative-peak GD
response at the latter, as seen in Figure 2.5b. A negative GD does not
mean that an input pulse to the ADMR will arrive at the through
port before it enters the input port, implying that causality is vio-
lated. Due to the strong GD dispersion (GDD) and attenuation near
a resonant frequency, the transmitted pulse at the through port is
so much distorted that its arrival time can no longer be unambig-
uously determined, for example, by tracking the pulse peak. The
interpretation of the GD as the transport time of the pulse energy
is no longer accurate and, thus, a negative GD value should not be
construed as a violation of causality (Brillouin 1960).
The GD at the drop port reaches a positive peak value given
by Sp,maxTrt. From Equation 2.57 we find that this value increases
for microrings with low loss and weak coupling, which is expected
since light in a high-Q microring is trapped for a longer period of
time before coupling out onto the bus waveguides. Since the GD
response at the drop port has the same shape as the power spectral
response, the GD bandwidth is the same as the transmission band-
width. There is, however, a tradeoff between this bandwidth and
the peak GD value. The GD–bandwidth product, GD × BW, which
is approximately the area underneath the GD response curve over
one FSR, is given by


GD × BW =

FSR
τ g,d dω = −

FSR

dω = −
∫ dψ.
FSR
(2.65)

Since the phase at the drop port changes by at most −π over one
FSR (see Figure 2.5a), the above equation implies that the GD–band-
width product of an ADMR is constrained by π. Thus any increase
in the peak GD must be offset by a corresponding decrease in the
bandwidth. The GD–bandwidth trade-off is an inherent limitation
in all optical delay elements based on passive optical resonators.
In Figure 2.6a and b we plot the phase and GD of an APMR for
different coupling values, showing the device responses in the under-
coupling, overcoupling, and critical coupling regimes. In all three
74 Optical Microring Resonators
(a) 0.5π (b) 30
τ = 0.95 Overcoupled
Critically coupled 20 τ = 0.9
0 Critically coupled

GD enhancement (S)
τ = 0.8
Phase response (ψ)

τ = 0.97 10 τ = 0.95 τ = 0.7


–0.5π τ = 0.98
τ = 0.99 0
Undercoupled
Undercoupled
–π Overcoupled –10 τ = 0.99
τ = 0.7
τ = 0.8 –20 τ = 0.98
–1.5π τ = 0.9
–30 τ = 0.97

–2π –40
–0.2π –0.1π 0 0.1π 0.2π –0.2π –0.1π 0 0.1π 0.2π
Roundtrip phase detune (∆φrt) Roundtrip phase detune (∆φrt)

Figure 2.6 (a) Phase response and (b) group delay response (S = τg/Trt) of
an APMR with various coupling coefficients and fixed round-trip attenu-
ation art = 0.95.

cases the phase of the output signal undergoes a sharp transition


at the resonances. However, in the overcoupling regime, the phase
transition slope is positive and becomes steeper for smaller couplings
(larger τ values). In the undercoupling regime, the phase has the oppo-
site behavior, exhibiting a positive slope which decreases for increas-
ing τ values. As a result, the GD of an APMR has a positive-peak
response for the overcoupling case and a negative-peak response for
the undercoupling case. At critical coupling, the phase undergoes a
discontinuous jump of π at resonance, which gives an infinite GD
value. Note that, in the overcoupling regimes, the phase of the output
signal changes by 2π over one FSR, so the GD–bandwidth product of
an APMR is constrained by 2π according to Equation 2.65.
The large GD enhancement in a microring resonator makes it
useful for realizing optical delay elements or buffers. However, as
mentioned above, the GD–bandwidth product places a limit on the
achievable peak GD for a given bandwidth. To further increase the
GD, we must increase the number of resonators in the device. The
analysis and design of arrays of microrings will be addressed in
Chapter 3. Another issue with microring delay elements is the GDD,
which must be kept small over the device bandwidth to minimize
signal distortion. The GDD is defined as

dτ g d2 ψ
Dg = ≈ −Trt2 2 , (2.66)
dω dφrt

where the approximation applies if we neglect GDD of the micro­


ring waveguide (i.e., the frequency dependence of Trt). Figure 2.7
Analytical Models of a Microring Resonator 75
600

Normalized GD dispersion
400 Overcoupled Undercoupled
τ = 0.9 τ = 0.99
200 τ = 0.85 τ = 0.98

–200

–400

–600
–0.1π –0.05π 0 0.05π 0.1π
Roundtrip phase detune (∆φrt)

Figure 2.7 Normalized GD dispersion, Dg /Trt2 of an APMR in the under-


coupling and overcoupling regimes. The round-trip attenuation of the
microring is fixed at art = 0.95.

shows the GDD of an APMR in the undercoupling and overcoupling


regime. In both cases the GDD is zero at resonance but can rise to a
large value over the device bandwidth. Note that the GDD curve has
a nearly linear slope at resonance, and the slope is ­positive for under-
coupled microrings and negative for overcoupled devices. Thus, by
tuning the coupling coefficient of an APMR, we can vary the GDD
slope over a wide range of values in both positive and negative direc-
tions. This behavior has been exploited to realized dispersion com-
pensators based on microring resonators (Lenz and Madsen 1999).

2.3 Energy Coupling Description


of a Microring Resonator
In the previous section, we derived the transfer functions of micro­
ring resonators in the add-drop and all-pass configurations using
the power coupling approach. The power coupling formalism is rig-
orous and general in that no assumption is made about the ­coupling
strengths, as long as the coupling junctions can be modeled by
lumped coupling elements.* In particular, the formalism can be used
to analyze strongly coupled microring resonators, or equivalently,
broadband microring devices. Power coupling analysis shows that

* The power coupling formalism is also valid for microring resonator devices
with complex coupling coefficients. For very strong coupling, the transmission
coefficient (τ) of the coupling junction will in general be complex, as can be seen
from Equation 1.128. Complex coupling coefficients also arise in certain micror-
ing coupling topologies, an example of which will be discussed in Section 3.7.
76 Optical Microring Resonators

the resonance spectrum of a microring resonator is periodic and


has the shape of an Airy function.
An alternative description of microring resonators that is valid
for weakly coupled or narrowband devices is the energy coupling
formalism, also known as “coupling of modes in time” formalism
(Haus 1984, Little et al. 1997). Under the assumption of weak cou-
pling, the field amplitude distribution around the microring can be
considered to be uniform. Neglecting the spatial dependence of the
field ­circulating around the microring, we can describe the behav-
ior of the device from the point of view of energy exchange between
the resonator and the external waveguides. Specifically, couplings
between the microring and bus waveguides are described by energy
coupling coefficients, which give the rates of energy coupled into and
out of the resonator. In effect, the microring is treated as a ­classical
Lorentz oscillator and, indeed, one can construct an equivalent LC
circuit of a microring resonator whose resonant ­frequency is equal
to ω 0 = LC (Van 2006). By neglecting the spatial ­distribution of the
field in the microring, the energy coupling ­formalism provides a
simpler model for analyzing microring devices than the power cou-
pling approach.
We consider first an uncoupled microring resonator with
­resonant frequency ω0. In the absence of any external coupling, the
only source of loss in the resonator is the propagation loss of the
microring waveguide. Using Equation 2.8, we can define the decay
rate of the mode amplitude due to intrinsic cavity loss as γ0 = vgα/2,*
where α and vg are the propagation loss and group velocity, respec-
tively, of the microring waveguide. According to Equation 2.7,
we can express the resonant mode in the uncoupled resonator as

a(t) = a0 e jω0t e − γ 0t , (2.67)

where a0 is the mode amplitude normalized so that |a0|2 gives the


initial stored energy in the microring. By differentiating Equation
2.67 with respect to time, we obtain the equation governing the time
decay of mode a(t):

da
= ( jω 0 − γ 0 )a. (2.68)
dt

* To account for the dispersion in the effective index of the microring waveguide,
we have replaced the effective index nr in Equation 2.8 by the group index ng.
Analytical Models of a Microring Resonator 77
si st
Input port Through port
(κ1) µ1, γ1

a(t)

γ0

(κ2) µ2, γ2
Drop port Add port
sd

Figure 2.8 Schematic of an ADMR with intrinsic decay rate γ0, input
energy coupling coefficient μ1 (field coupling κ1), and output energy cou-
pling coefficient μ2 (field coupling κ2).

Consider now a microring resonator coupled to an input


waveguide and an output waveguide with input and output field
­coupling coefficients κ1 and κ2, respectively, as shown in Figure 2.8.
The external couplings lead to additional losses in the microring
due to energy transfer to the two waveguides. Denoting the ampli-
tude decay rates due to coupling to the input and output bus wave-
guides as γ1 and γ2, respectively, we modify Equation 2.68 to read

da
= ( jω 0 − γ 0 − γ 1 − γ 2 )a. (2.69)
dt

In addition, suppose the input waveguide carries a signal si toward


the resonator, with the amplitude normalized such that |si|2 repre-
sents the power flowing in the input waveguide. The input signal
supplies energy to the resonator at a rate given by −jμ1si, where μ1 is
the input energy coupling coefficient and the phase –j is c­ hosen for
consistency with the coupling of modes in space formalism (i.e., the
term −jκ1si in Equation 2.15). Adding this term to Equation 2.69, we
obtain the equation governing the mode amplitude a(t) in an ADMR:

da
= ( jω 0 − γ 0 − γ 1 − γ 2 )a − jµ1si . (2.70)
dt

We can determine the relationship between the energy cou-


pling coefficient μ1 and the decay rate γ1 from power conservation
consideration (Little et al. 1997). Specifically, we consider a micror-
ing resonator with no intrinsic loss (γ0 = 0) which is coupled to only
78 Optical Microring Resonators

an input waveguide (γ2 = 0). Suppose the microring is charged to an


initial energy |a0|2 by the input signal si. After the input signal is
removed (si = 0), the subsequent energy decay is given by
2 2
a(t) = a0 e −2 γ 1t . (2.71)

The rate of energy decay from the microring is thus


2
d a(t) 2
(2.72)
= −2γ 1 a(t) .
dt

The energy leaving the microring is coupled to the input waveguide


and is transmitted to the through port, so we must have |st|2 = −
d|a(t)|2/dt. Since the power transmitted at the through port is also
equal to |st|2 = µ12|a(t)|2, we obtain the relation

µ12 = 2γ 1 . (2.73)

Similarly, at the output coupling junction of an ADMR, the energy


coupling coefficient μ2 is related to the output decay rate γ2 by
µ 22 = 2γ 2.
It is also useful to derive expressions relating the energy
­coupling coefficients μ1 and μ2 to the field coupling coefficients κ1
and κ2. First we note that the energy in the microring is related to
the power flow by

2 2 2 2πR
a(t) = A(t) Trt = A(t) , (2.74)
vg

where A(t) is the amplitude of the wave traveling in the micror-


ing normalized such that |A(t)|2 gives the power flow through any
cross section of the microring waveguide at time t. In the absence of
an input signal, the power coupled from the microring to the input
waveguide is
2 2 2
st = κ 12 A(t) = µ12 a(t) . (2.75)

Using the relationship between |a(t)|2 and |A(t)|2 in Equation 2.74,


we obtain

2πR 2
κ 12 = µ12Trt = µ1 . (2.76)
vg
Analytical Models of a Microring Resonator 79

Since Trt = 1/Δf FSR, we can also write κ 12 = µ12 /∆f FSR. Similarly, the
output field coupling coefficient is related to the output energy
­coupling coefficient by κ 22 = µ 22Trt = µ 22 /∆f FSR.
We now solve Equation 2.70 for the case when the microring
is excited by a monochromatic wave of frequency ω, si ~ ejωt. Since
the field in the microring will also have the same harmonic time
­dependence, a ~ ejωt, we obtain from Equation 2.70

jωa = ( jω 0 − γ )a − jµ1si , (2.77)

where γ = γ0 + γ1 + γ2 is the total decay rate. Solving the above equa-


tion for a, we get

− jµ1si
a= , (2.78)
j(ω − ω 0 ) + γ

which gives the frequency response of the energy amplitude in


the microring. Since sd = −jμ2 a and st = si − jμ1a, we obtain the trans-
fer functions at the drop port and through port of the ADMR as
follows:

sd − µ 1µ 2
= , (2.79)
si j(ω − ω 0 ) + γ

st j(ω − ω 0 ) + γ − µ12
= . (2.80)
si j(ω − ω 0 ) + γ

Defining the complex frequency variable, s = j(ω − ω0), we can write


the above expressions as

− µ 1µ 2 − µ 1µ 2
Hd (s) = = , (2.81)
s+ γ s + γ 0 + µ12 /2 + µ 22 /2

s + γ − µ12 s + γ 0 − µ12 /2 + µ 22 /2
H t (s) = = . (2.82)
s+ γ s + γ 0 + µ12 /2 + µ 22 /2

The above expressions have the form of a transfer function of an


analog filter* with a single pole at s = −γ. We also note that the drop

* The inverse Laplace transforms of Hd(s) and Ht(s) give the impulse responses at
the drop port and through port, respectively, of the ADMR.
80 Optical Microring Resonators

port transfer function has no zero, whereas the through port trans-
fer function has a zero at s = µ12 − γ .
The power spectral responses of the ADMR are obtained by
taking the absolute square of Equations 2.79 and 2.80:

2
s µ12 µ 22
Td (ω) = d = , (2.83)
si (ω − ω 0 )2 + γ 2

2
st (ω − ω 0 )2 + ( γ − µ12 )2
Tt (ω) = = . (2.84)
si (ω − ω 0 )2 + γ 2

Equation 2.83 indicates that the drop port response has the shape
of a Lorentzian resonance centered at the resonant frequency
ω0. The FWHM bandwidth is given by ΔωBW = 2γ, or in terms of
wavelength:

λ 02 γ
∆λ BW = . (2.85)
πc

To determine the Q factor of the ADMR, we can use the


e­ xpression Q = λ0/ΔλBW to get Q = πc/λ0γ = ω0/2γ. The same for-
mula can also be obtained directly from the definition of the
Q factor:

Stored energy
Q = ω0 . (2.86)
Power loss

Since the stored energy in the microring is |a|2 and the total power
loss is 2γ|a|2, we have

|a|2 ω
Q = ω0 = 0. (2.87)
2γ |a| 2γ
2

If we define the Q factor due to intrinsic loss as Q0 = ω0/2γ0, and


the Q factor due to coupling to each of the bus waveguides as
Q1 = ω 0 /2γ 1 = ω 0 /µ12 and Q2 = ω 0 /2γ 2 = ω 0 /µ 22 ; then, from Equation
2.87, we have

1 1 1 1
= + + . (2.88)
Q Q0 Q1 Q2
Analytical Models of a Microring Resonator 81

For an ADMR with low loss (γ0 ≈ 0) and symmetric coupling


(μ1 = μ2 = μ), we have Q0 → ∞, Q1 = Q2 = ω0/μ2, so

ω0 2π 2 ng R
Q= = , (2.89)
2µ 2 λ 0κ 2

which agrees with the expression for the loaded Q factor in


Equation 2.40.
For an APMR with energy coupling coefficient μ, the transfer
function can be obtained by setting μ1 = μ and μ2 = 0 in Equation 2.82

s + γ − µ 2 s + γ 0 − µ 2 /2
H ap (s) = = . (2.90)
s+ γ s + γ 0 + µ 2 /2

From the above equation we see that complete extinction of the


transmitted signal is achieved at resonance if the decay rate due to
external coupling, γe = μ2/2, is equal to the decay rate due to intrin-
sic loss, γe = γ0. This is the critical coupling condition. The APMR is
undercoupled if γe < γ0 and overcoupled if γe > γ0.
We can also obtain the phase and GD responses of an ADMR
from its transfer functions in Equations 2.81 and 2.82. The phase
responses at the drop port and through port are given by

 ω − ω0 
ψ d (ω) = π − tan −1  , (2.91)
 γ 

 ω − ω0   ω − ω0 
ψ t (ω) = tan −1  − tan −1  . (2.92)
2 
 γ − µ1   γ 

The GD responses are obtained by differentiating the phase with


respect to ω. The results for the GD at the drop port and through
port are

γ
τ g,d (ω) = , (2.93)
(ω − ω 0 )2 + γ 2

γ − µ12 γ
τ g,t (ω) = − 2 2 2
+ . (2.94)
(ω − ω 0 ) + ( γ − µ1 ) (ω − ω 0 )2 + γ 2

From Equation 2.93 we find that the peak GD at the drop port is
τg,max = 1/γ, which is equal to the decay time of the wave amplitude in
82 Optical Microring Resonators

the microring. The FWHM bandwidth of the GD response at the drop


port is ΔωBW = 2γ, so the GD–bandwidth product is equal to 2. Finally,
we note that the phase and GD responses of an APMR are also given
by Equations 2.92 and 2.94, respectively, with μ1 = μ and μ2 = 0.

2.4 Relationship between Energy Coupling


and Power Coupling Formalisms
Further insight into the energy coupling formalism can be gained
by establishing a direct relationship between the power coupling
and energy coupling formalisms. In particular, we will show that
in the limit of low loss and weak coupling, the power coupling for-
malism reduces to the energy coupling formalism. In Section 2.2,
we found that the power-normalized field inside an ADMR is given
by (Equation 2.21):

− jκ 1si
A= . (2.95)
1 − τ1τ 2 art e − jφrt

Writing the round-trip phase in the microring as ϕrt = 2mπ + Δϕrt,


where Δϕrt is the phase detune from the resonance, we have

e − jφrt = e − j∆φrt = e − j∆ωTrt , (2.96)

where Δω = ω − ω0 and Trt = 2πR/vg is the round-trip time of the


microring. The round-trip amplitude attenuation factor art can
­likewise be written as

art = e − απR = e − γ 0Trt , (2.97)

where γ0 = αvg/2 is the intrinsic decay rate due to loss in the micror-
ing. The denominator of Equation 2.95 can now be expressed as

1 − τ1τ 2 art e − jφrt = 1 − exp[ln(τ1 ) + ln(τ 2 ) − ( γ 0 + j∆ω)Trt ]. (2.98)

For weak couplings we have

−κ 2 µ2
ln(τ1 ) =
1
2
( 2
)
ln 1 − κ 12 ≈ 1 = − 1 Trt ,
2
(2.99)

with a similar expression for ln(τ2). With these approximations,


Equation 2.98 becomes
Analytical Models of a Microring Resonator 83

 µ2 µ2 
1 − τ1τ 2 art e − jφrt ≈  1 + 2 + ( γ 0 + j∆ω) Trt , (2.100)
 2 2 

where we have also used the small argument approximation for


the exponential function, which is valid for weak couplings, small
loss and small frequency detunes around the microring resonance.
Substituting Equation 2.100 into Equation 2.95, we get

− j(µ1 Trt )si


A≈ . (2.101)
[µ /2 + µ 22 /2 − ( γ 0 + j∆ω)]Trt
2
1

Since the energy amplitude a in the microring is related to the power-


normalized field A by a = A Trt , we obtain from Equation 2.101

− jµ1si
a= . (2.102)
( j∆ω + γ 0 ) + µ12 /2 + µ 22 /2

The above expression for the energy amplitude in the microring


is the same as Equation 2.78 obtained using the energy coupling
formalism. We thus conclude that the energy coupling formalism is
an approximation of the power coupling formalism under the weak
coupling and low-loss (or equivalently, narrowband) condition.
Figure 2.9 compares the spectral responses of a relatively
strongly coupled ADMR obtained using the power coupling and

–5
Transmission (dB)

Td
–10

–15
Tt
–20

–25
–π 0 π 2π
Roundtrip phase detune (∆φrt)

Figure 2.9 Power spectral responses at the drop port (Td) and through
port (Tt) of an ADMR with power coupling coefficients κ 12 = κ 22 = 0.2 and
5% round-trip power loss (art = 0.975). Solid lines are spectra obtained
using the power coupling formalism; dashed lines are spectra obtained
using the energy coupling formalism.
84 Optical Microring Resonators

energy coupling approaches. The power coupling coefficients are


κ 12 = κ 22 = 0.2 and the round-trip power loss is 5%. One notable
difference between the two results is that the spectral responses
obtained from power coupling are periodic, whereas the energy
coupling formalism can model only one resonance (the center res-
onance at Δϕrt = 0). In addition, we observe that the discrepancy
between the two sets of spectra becomes more pronounced as we
move away from the center resonance. Thus the energy ­coupling
approximation is only good for frequencies near a microring
resonance.

2.5 Summary
In this chapter, we developed two alternate analytical models for a
microring resonator: a rigorous model based on power coupling and
an approximate but simpler model based on energy coupling. Using
these models, we obtained the transfer functions of the microring
in the add-drop and all-pass configurations, and derived useful
formulas for determining important quantities characterizing the
performance of these devices. In subsequent chapters, we will also
make extensive use of both the power coupling and energy coupling
formalisms to analyze and design more complex device configu-
rations and more advanced applications of microring ­resonators.
More specifically, we will extend the analysis techniques developed
in this chapter to design optical filters based on coupled microring
resonators in Chapter 3. The behaviors of microring resonators in
the presence of optically-induced and electrically-induced material
nonlinearities will be studied in Chapters 4 and 5, respectively.

References
Brillouin, L. 1960. Wave Propagation and Group Velocity. New York:
Academic Press.
Haus, H. A. 1984. Waves and Fields in Optoelectronics. Englewood
Cliffs, NJ: Prentice-Hall.
Lenz, G., Madsen, C. K. 1999. General optical all-pass filter
­structures for dispersion control in WDM systems. J. Lightwave
Technol. 17(7): 1248–1254.
Analytical Models of a Microring Resonator 85

Little, B. E., Chu, S. T., Haus, H. A., Foresi, J., Laine, J.-P. 1997. Microring
resonator channel dropping filters. J. Lightwave Technol. 15(6):
998–1005.
Madsen, C. K., Zhao, J. H. 1999. Optical Filter Design and Analysis:
A Signal Processing Approach. New York: John Wiley & Sons Inc.
Van, V. 2006. Circuit-based method for synthesizing serially
­coupled microring filters. J. Lightwave Technol. 24(7): 2912–2919.
Yariv, A. 2000. Universal relations for coupling of optical power
between microresonators and dielectric waveguides. Electron.
Lett. 36(4): 321–322.
Chapter 3

Coupled Microring Optical Filters

One of the most important applications of microring resonators


is the synthesis of optical filters. In the broadest sense, a filter is
a structure or device that alters the amplitude, phase, or group-
delay characteristics of an input signal in a desired fashion. Owing
to their compact sizes and strongly dispersive characteristics near
a resonant frequency, microring resonators have been shown to
be ­ particularly versatile elements for constructing high-order
­integrated optical filters.
This chapter introduces the theories and techniques for ana-
lyzing and designing coupled microring resonator (CMR) optical
filters. We will begin in Section 3.1 with a discussion of infinite
arrays of microring resonators and show how the periodicity of
these structures gives rise to frequency responses with passbands
and stopbands. We will then turn our focus to finite arrays of
CMRs and study how their spectral characteristics can be tailored
to achieve a desired filter response. Representations of microring
filters as analog and digital IIR filters are discussed in Section 3.2
along with a review of the general forms of the transfer functions
of each type of filters. The remainder of the chapter will be devoted
to the analysis and design of specific microring filter configura-
tions. Although numerous variations of microring filter designs
have been proposed and studied in the literature, we will restrict
our treatment to the more universal types, namely, cascaded all-
pass and add-drop microring filters, and serially coupled micror-
ing ­filters and their generalizations to 2D coupling topologies.
With the exception of the cascaded add-drop microring array, each
of these structures can exactly realize a specific type of transfer
function of a given order. For each type of filters, we will develop
techniques for analyzing and designing high-order optical filters
based on the power coupling formalism. Simpler techniques based
on energy coupling will also be developed for coupled microring
networks of 1D and 2D coupling topologies.

87
88 Optical Microring Resonators

3.1 Periodic Arrays of Microring Resonators


Periodic microring arrays of infinite length behave as 1D artificial
media with distributed resonances. These structures have unique
photonic band structures which give rise to many interesting light
propagation characteristics. Three particular configurations—the
all-pass microring array, the cascaded add-drop microring array,
and the serially coupled microring array—have been well stud-
ied. The serially coupled microring array is also known as CROW
(Coupled Resonator Optical Waveguides) (Yariv et al. 1999), while
parallel cascaded arrays of all-pass and add-drop microrings are
sometimes called single-channel and double-channel SCISSORs (for
side-coupled integrated spaced sequence of resonators) (Heebner
et al. 2004), respectively. The double-channel SCISSOR and CROW
structures share the common feature that the distributed resonances
and feedback give rise to photonic band structures consisting of
alternating transmission bands and ­stopbands. Strong dispersion
occurs near the band edges, giving rise to slow light effects which
can be exploited for applications in optical delay lines and group
velocity dispersion engineering (Melloni et al. 2003).
In general, the dispersion characteristics of a periodic array
of distributed resonators can be analyzed using the Bloch matrix
­formalism (Yeh et al. 1977, Heebner et al. 2004). The array can be
modeled by a periodic sequence of unit elements as shown in
Figure 3.1, with each element described by a transfer matrix (or
transmission matrix) T defined as
 ak +1  T11 T12   ak 
 b  = T T22   bk 
. (3.1)
 k +1   21
For an infinite array of identical elements with spatial periodicity Λ,
Bloch’s theorem states that the output fields and input fields of each
unit element are related by
k–1 Element k k+1
ak ak+1

T T T

bk bk+1
Λ

Figure 3.1 Schematic of an infinite periodic array of identical unit


­elements, each described by transfer matrix T.
Coupled Microring Optical Filters 89

 ak +1   ak  − jβΛ
b  = b  e , (3.2)
 k +1   k 

where β is the effective propagation constant or Bloch vector.


If there is loss in the array, the propagation constant will be com-
plex. Upon substituting Equation 3.2 into Equation 3.1, we obtain

T11 T12   ak   ak  − jβΛ


T = e , (3.3)
 21 T22   bk   bk 

which implies that λ = e−jβΛ are the eigenvalues of the matrix T.


These eigenvalues are the roots of the characteristic equation

∆ T e jβΛ + e − jβΛ = T11 + T22 , (3.4)

where ΔT = T11T22 − T12T21 is the determinant of the transfer matrix.


For a lossless unit element, the matrix T is unitary, that is, its deter-
minant is equal to 1. In this case Equation 3.4 can be simplified to
read

1
cos(βΛ) = (T11 + T22 ). (3.5)
2

In addition, for symmetrical networks we have T22 = T11∗ , in which


case Equation 3.5 further reduces to

cos(βΛ) = Re {T11 } . (3.6)

Equation 3.6 gives the dispersion relation of a periodic array of


unit elements characterized by transfer matrix T. Note that the
dispersion relation depends only on the parameters T11 and T22.
In the following subsections, we examine in detail the transmis-
sion ­characteristics of several common types of periodic microring
arrays.

3.1.1 Periodic arrays of APMRs


An all-pass microring array consists of an infinite sequence of iden-
tical APMRs, as depicted in Figure 3.2. Each microring has radius
R and is coupled to a common bus waveguide with field coupling
coefficient κ (transmission coefficient τ). The microring waveguide is
90 Optical Microring Resonators

k−1 k k+1
κ κ κ
Λ Λ
ak ck ak+1 ck+1

Figure 3.2 Schematic of a periodic array of APMRs.

assumed to have effective index nr and propagation loss α. The out-


put of each microring is fed forward to the next stage, so there is no
feedback in the array. The absence of distributed feedback implies
that the structure has no Bragg bandgap. Each unit element in the
array consists of an APMR and a bus waveguide of length Λ. The
output of element k is given by ak+1 = T11ak, where T11 is the ­product
of the transfer function of the APMR and a phase delay e−jθ due to
the bus waveguide:

τ − art e − jφ − jθ
T11 = H ap e − jθ = − jφ
e ≡ e − jΦ . (3.7)
1 − τart e

In the above equation, ϕ = 2πnrkR and art = exp(−παR) are the round-
trip phase and amplitude attenuation, respectively, of the micror-
ing. The angle θ is given by θ = nbkΛ, where nb is the effective index
of the bus waveguide. The phase angle Φ of T11 defined in Equation
3.7 is real if there is no loss in the microrings (|Hap| = 1); otherwise
it will be complex. Applying Bloch’s ­theorem, we have

ak +1 = ak e − jΦ = ak e − jβΛ , (3.8)

which gives βΛ = Φ.
For a lossless APMR (art = 1), the transfer function can be
expressed as

τ − e − jφ − jφ 1 − τe

H ap = = − e , (3.9)
1 − τe − jφ 1 − τe − jφ

from which we obtain the phase angle of Hap as follows:

 sin φ 
ψ ap = π − φ − 2 tan −1  . (3.10)
 τ − cos φ 
Coupled Microring Optical Filters 91

The dispersion relation for the Bloch wave vector of the APMR
array is thus βΛ = Φ = θ − ψap, or

1 −1  sin φ  
β= φ + 2 tan   − π + θ . (3.11)
Λ  τ − cos φ  

If there is loss in the microrings, we can obtain the phase angle ψap
from the more general expression in Equation 2.59 with z0 = τ/art
and p0 = 1/τart. The dispersion relation of the APMR array is given
by Re{βΛ} = θ − ψap, or

1 −1  τart sin φ   a sin φ  


Re{β} =  tan   − tan −1  rt  + θ . (3.12)
Λ  1 − τart cos φ   τ − art cos φ  

Note that except for an additional phase shift θ due to the bus wave-
guide, the phase response per period of the APMR array is the same
as that of a single APMR.
Figure 3.3a and b compare the dispersion diagram (plot of ϕ vs.
βΛ) of an array of overcoupled microrings with that of an array of
undercoupled microrings. In both arrays the length Λ of the bus
waveguide is chosen to be equal to the microring circumference
so that we have θ = ϕ (assuming that nr = nb). We will also assume
nr and nb to have negligible dispersion so that the group index of
the waveguide is the same as the effective index. The overcoupled
microrings have transmission coefficient τ = 0.9 and round-trip
attenuation art = 0.99, whereas the parameters for the undercoupled
microrings are τ = 0.9 and art = 0.85. Also shown in the figures are
the normalized group index, ng/nr = Λ(dβ/dϕ), and the power trans-
mission through 10 periods of the array. The transmission is com-
puted from |T11|2N, where N = 10 is the number of periods. From the
plots we observe that the dispersion characteristics of the arrays
are dominated by the resonances of the all-pass microrings, with
the group index exhibiting strong dispersion near each microring
resonance. For the array of overcoupled AMPRs, the group index is
positive and becomes significantly larger than the effective index
of the microring waveguide near each resonance, indicating that
light is slowed down as it propagates through the array. For the
array of undercoupled APMRs, the group index becomes negative
over a narrow band of frequencies centered around each resonance.
This is the regime of fast light, or superluminal propagation, where
92 Optical Microring Resonators
(a)
3π 3π 3π
Microring roundtrip phase, φ 2.5π 2.5π 2.5π
2π 2π 2π
1.5π 1.5π 1.5π
π π π
0.5π 0.5π 0.5π
0 0 0
–0.5π –0.5π –0.5π
–π –π –π
–π –0.5π 0 0.5π π 0 2 4 6 0 0.5 1
βΛ ng/nr Power transmission
(b)
3π 3π 3π
Microring roundtrip phase, φ

2.5π 2.5π 2.5π


2π 2π 2π
1.5π 1.5π 1.5π
π π π
0.5π 0.5π 0.5π
0 0 0
–0.5π –0.5π –0.5π
–π –π –π
–π –0.5π 0 0.5π π –4 –2 0 0 0.5 1
βΛ ng/nr Power transmission

Figure 3.3 Dispersion characteristics of APMR arrays with Λ = 2πR:


(a) array of overcoupled microrings (τ = 0.9; art = 0.99), (b) array of under-
coupled microrings (τ = 0.9; art = 0.85). Left panel: dispersion diagram;
­center panel: normalized group index ng/nr; right panel: power transmis-
sion through 10 periods. (After Heebner, J. E., et al., 2004, J. Opt. Soc. Am. B
21(10): 1818–1832.)

a light pulse can appear to travel backward or take less time to


traverse a section of the array than in a straight waveguide of the
same length. However, the transmitted light signal is also severely
attenuated and distorted due to the strong group-delay dispersion
in the stopband.
Note that the dispersion diagrams of both APMR arrays do
not show any bandgap since there is no feedback in the arrays.
For an ideal lossless APMR array, the transmission is unity for
all frequencies. However, in practical structures, a slight amount
of microring loss can cause sharp transmission dips to occur
around the resonances, as evidenced in the transmission plot of
the overcoupled microring array in Figure 3.3a. As the micror-
ing loss increases, the stopband widens further, as seen for the
case of the undercoupled array in Figure 3.3b. To access the slow
light (or fast light) regime, one would have to operate within the
Coupled Microring Optical Filters 93

stopband, which results in significant attenuation of the signal.


This is a major drawback of APMR arrays for slow light applica-
tions. Nevertheless, an array of 56 APMRs has been experimen-
tally demonstrated in the SOI material system for on-chip optical
buffer applications (Xia et al. 2007b).

3.1.2 Periodic arrays of ADMRs


Figure 3.4 shows a schematic of a periodic array of parallel cascaded
ADMRs. We assume the ADMRs to be identical with equal input
and output field coupling coefficients κ (transmission coefficients τ).
Adjacent ADMRs are connected via two parallel input and output
bus waveguides of length Λ, which is assumed to be greater than
the microring diameter so that there is no direct coupling between
two adjacent microrings. A wave traveling in the input bus wave-
guide will be partially coupled to each microring and subsequently
transmitted to the output bus waveguide, where it propagates
backward as a reflected wave. The array can thus be regarded as a
periodic grating in which the microrings act as reflecting elements.
Therefore, we expect the structure to have Bragg bandgaps centered
at wavelengths satisfying the condition 2Λ = mB(λ/nb), where mB is
an integer and nb is the effective index of the bus waveguides. In
addition, since the ADMRs are strongly reflecting at wavelengths
corresponding to the microring resonances, there are also stop-
bands occurring at wavelengths satisfying the resonance condition
2πR = mR(λ/nr), where mR is an integer and nr is the effective index
of the microring waveguide. Overlapping between the Bragg band-
gaps and microring stopbands occurs if the ratio Λ/πR is equal to
nrmB/nbmR.

ak ck ak+1 ck+1
Λ Λ Transmitted wave
Input bus κ κ κ

k−1 k k+1
Output bus κ κ κ
Λ Λ
Reflected wave bk dk bk+1 dk+1

Figure 3.4 Schematic of a periodic array of cascaded add-drop microring


resonators.
94 Optical Microring Resonators

Each unit element in the array consists of an ADMR and two


parallel waveguide segments of length Λ. For simplicity, we assume
that the microrings are lossless. Using Equations 2.23 and 2.24 for
the drop port and through port transfer functions of an ADMR,
we can write, for each microring k,

bk  1  −κ 2 e − jφ/2 τ(1 − e − jφ )  ak 


c  = 2 − jφ   , (3.13)
 k  1− τ e −κ 2 e − jφ/2   dk 
− jφ
 τ(1 − e )

where ϕ is the microring round-trip phase. The above matrix


­equation can be recast in the form of a transfer matrix Tad of the
ADMR as

 ck  1 (τ 2 − e − jφ ) −κ 2 e − jφ/2   ak   ak 
d  = − jφ     ≡ Tad b  . (3.14)
 k  τ(1 − e )  κ e (1 − τ 2 e − jφ )  bk 
2 − jφ/2
 k

The transfer matrix of the two parallel bus waveguides of length Λ


is given by

 ak + 1   e 0   ck 
− jθ
 ck 
b  =  
jθ  d 
≡ Λ , (3.15)
 k + 1   0 e   k   dk 

where θ = nbkΛ. From Equations 3.14 and 3.15, we obtain the trans-
fer matrix of each unit element in the array as T = ΛTad, which is a
­unitary matrix with the elements T11 and T22 given by

τ 2 − e − jφ − jθ
T11 = e = T22∗ . (3.16)
τ(1 − e − jφ )

Substituting the above expression for T11 into Equation 3.6, we arrive
at the following dispersion relation for the ADMR array:

1  1 τ φ  1  φ  
β=± cos −1   sin  + θ + sin  − θ   . (3.17)
Λ  sin(φ/2)  2  2  2τ  2  

The plus and minus signs correspond to the forward- and


­backward-propagating Bloch waves, respectively, in the array.
Figure 3.5 shows the dispersion diagram (plot of ϕ vs. βΛ) of
an array of lossless ADMRs with τ = 0.8, Λ/2πR = 2/3, and effective
index nr = nb, which is assumed to be dispersionless. The normalized
Coupled Microring Optical Filters 95
7π 7π 7π
Microring roundtrip phase, φ 6π 6π 6π
5π 5π 5π
4π 4π 4π
3π Bragg bandgap 3π 3π
2π Resonator bandgap 2π 2π
Bragg bandgap
π π π
0 Bragg + resonator bandgap 0 0
−π −π −π
–1 –0.5 0 0.5 1 0 0.5 1 0 0.5 1
βΛ vg/vr Power transmission

Figure 3.5 Dispersion characteristics of a lossless ADMR array with


τ = 0.8 and Λ/2πR = 2/3. Left panel: dispersion diagram; center panel: nor-
malized group velocity (vg/vr); right panel: power transmission through a
10-period array. (After Heebner, J. E., et al., 2004, J. Opt. Soc. Am. B 21(10):
1818–1832.)

group velocity, vg/vr where vr = c/nr, and the power transmission*


through 10 periods are also shown. The dispersion diagram dis-
plays a complex pattern consisting of stopbands due to both dis-
tributed feedback and microring resonances. Note that the Bragg
stopbands are of the direct bandgap type since the band m ­ axima
align with the band minima, whereas those due to the microring
resonances are of the indirect type (Heebner et al. 2004). The two
types of stopbands overlap every 6π change in the microring round-
trip phase, which corresponds to three free spectral ranges (FSRs)
of the resonator. From the plot of the normalized group velocity,
we observe that within each transmission band there is a reduction
in the group velocity of the Bloch mode compared to the group
velocity in the waveguide (vg < vr). This slow light effect is caused
by light being trapped in the microrings for long periods of time
(equal to the microring lifetime) and/or undergoing multiple Bragg
reflections as it propagates along the array. The group ­velocity can
be further reduced by decreasing the microring coupling strength
κ to increase the lifetime of each resonator.
The plot of the power transmission through the array shows
that within each passband, the transmission reaches unity but also
exhibits large ripples. These ripples are caused by the fact that the
array in the example is truncated with a finite number of periods

* Formulas for computing the transmission and reflection spectra of an ADMR


array of finite length are derived in Section 3.5.
96 Optical Microring Resonators
Unit cell

κ κ
ck–1 ak dk bk+1
dk–1 bk ck ak+1

k−1 k k+1

Figure 3.6 Schematic of a CROW lattice consisting of a periodic array of


serially coupled microring resonators.

(N = 10), resulting in reflections of the Bloch waves at the termina-


tions. The transmission spectra flatten out in the limit N becomes
very large. For an ADMR array of finite length, the ripples in the
transmission response can be reduced by applying apodization to
the coupling coefficients. We will look at the analysis and design of
this type of structures in more detail in Section 3.5.

3.1.3 Arrays of serially coupled microring resonators


Figure 3.6 shows a schematic of an infinite chain of serially coupled
microring resonators, which is also known as a Coupled Resonator
Optical Waveguide (CROW) (Yariv et al. 1999). The microrings are
assumed to be identical with radius R and adjacent microrings are
directly coupled to each other via the field coupling coefficient κ
(transmission coefficient τ). Since energy in a microring can be cou-
pled forward and backward to its two neighbors, the CROW array
supports both forward- and backward-propagating Bloch modes.
At resonance, energy transport along the array is considerably
slowed down since it takes time for light to “charge” each micror-
ing before moving to the next. At off-resonance wavelengths, light
is prevented from building up in the microrings and hence cannot
propagate in the array. Thus, we expect the frequency response of
the array to have stopbands separating the microring resonances.
Indeed, a CROW array can be regarded as a distributed feedback
grating in which each coupling junction acts as a partial reflector.
In this case, transmission bands occur when the Bragg condition
2Λ = m(λ/nr) is satisfied,* where Λ = πR is the length of each period.

* The Bragg condition 2Λ = mλ/nr corresponds to passbands rather than stop-


bands in this case because the transmission coefficient of the partial reflector is
imaginary (equal to −jκ).
Coupled Microring Optical Filters 97

Note that this Bragg condition is also the resonance condition of the
microrings.
To derive the dispersion relation of an infinite CROW array, we
define each unit element in the array as consisting of a microring
and a coupling junction, as shown in Figure 3.6 (Poon et al. 2004).
For simplicity, we will assume that the microrings have no loss.
Within microring k, the fields [ck, dk] are related to [ak, bk] by

 ck   0 e − jφ/2   ak   ak 
 d  =  jφ/2   ≡ Λ , (3.18)
 k   e 0   bk   bk 

where ϕ is the microring round-trip phase. At the coupling junction


between microrings k and k + 1, we have the relations

dk = τck − jκak +1 , (3.19)

bk +1 = τak +1 − jκck .
(3.20)
The above expressions can be rearranged to give the transfer matrix
K of the coupling junction as follows:

 ak + 1  1  τ −1  ck   ck 
 b  = jκ 1   
− τ   dk 
≡ K  . (3.21)
 k +1    dk 

The transfer matrix of each unit element in the CROW array can be
obtained using Equations 3.18 and 3.21:

1  −e τe − jφ/2 
jφ/2
Λ=
T = KΛ  . (3.22)
jκ  − τe jφ/2 e − jφ/2 
∗ jφ/2
The above matrix is unitary with T11 = T22 = − e /jκ . Using
Equation 3.6, we obtain the following dispersion relation for the
CROW array:

1  1 
β=± cos −1  − sin(φ/2) . (3.23)
Λ  κ 

The plus and minus signs correspond to the forward- and


­backward-propagating Bloch waves, respectively, in the array.
Figure 3.7 shows the dispersion diagram (plot of ϕ vs. βΛ) of a
lossless CROW array with τ = 0.8. We assume the effective index nr
98 Optical Microring Resonators
7π 7π 7π
Microring roundtrip phase, φ 6π 6π 6π
5π 5π 5π
4π 4π 4π
3π 3π 3π
2π 2π 2π
π π π
0 0 0
−π −π −π
–π –0.5π 0 0.5π π 0 0.2 0.4 0.6 0 0.5 1
βΛ vg/vr Power transmission

Figure 3.7 Dispersion characteristics of a lossless CROW with τ = 0.8.


Left panel: dispersion diagram; center panel: normalized group veloc-
ity (vg/vr); right panel: power transmission through five microrings.
(After Heebner, J. E., et al., 2004, J. Opt. Soc. Am. B 21(10): 1818–1832.)

of the microring waveguide to have negligible dispersion so that its


group index is also the same as the effective index. Also shown in
the figure are the normalized group velocity, vg/vr where vr = c/nr,
and the power transmission spectrum* through a CROW array con-
sisting of N = 5 microrings. The dispersion diagram shows the peri-
odic formation of a passband and a stopband for every 2π change in
the microring round-trip phase (or one microring FSR). In the plot
of the n ­ ormalized group velocity, we observe that within each pass-
band the group velocity is reduced (vg < vr) as light spends more
time in each microring before coupling to the next. In general, the
group ­velocity can be further reduced by decreasing the coupling
­coefficient κ between the microrings. For the lossless CROW array
of finite length, the power transmission reaches unity in the pass-
band although there are also large ripples. These ripples are caused
by reflections of the Bloch waves from the two terminated ends of
the CROW array. For a finite CROW array, the ripples in the trans-
mission response can be reduced by proper adjustment (or apodiza-
tion) of the coupling coefficients. The analysis and design of CROW
­filters will be treated in Section 3.4.
Experimentally, CROW arrays of 100 micro-racetracks (Xia
et al. 2007b) and 235 micro-racetracks (Cooper et al. 2010) have
been ­demonstrated in the SOI material system for on-chip buff-
ering applications. A major challenge in the realization of these
long CROW arrays is that due to fabrication imperfections, the

* The transmission response of a finite CROW array is derived in Section 3.4.


Coupled Microring Optical Filters 99

microring resonators do not have identical resonant frequencies.


This ­resonance mismatch causes high transmission loss and large
­ripples in the passband.
The analysis of 1D CROW arrays presented in this section can
also be extended to obtain the dispersion relation of a 2D lattice of
CMRs. In particular, it can be shown that the dispersion relation of
an infinite 2D microring lattice can be expressed as a sum of two
independent dispersion relations of 1D microring arrays in the two
orthogonal directions (Chremmos and Uzunoglu 2008). Of more
practical interest are the spectral responses of finite 2D microring
lattices. In Section 3.7 we will develop methods for ­analyzing and
designing microring filters of 2D coupling topologies.

3.2 Transfer Functions of Coupled


Microring Optical Filters
Infinite periodic arrays of microring resonators are idealized
structures since in practice the arrays must have finite lengths.
However, truncation of an infinite microring array gives rise to
surface states localized near the ends of the array, which mani-
fest themselves as ripples in the transmission response. These
ripples may also be understood as being caused by reflections due
to impedance ­mismatch between the Bloch modes of the array
and the excitation and transmission waves at the input and out-
put ports. We can reduce or eliminate this impedance mismatch
by apodizing the coupling coefficients in the array. The problem
of determining a set of coupling coefficients that can ­generate a
desired transmission response is called the microring filter syn-
thesis problem.
In this chapter, we approach the analysis and design of micror-
ing optical filters from the point of view of classical filter theory.
Specifically, the spectral response of a microring filter can be under-
stood in terms of the poles and zeros of its transfer functions, and
the design of the filter is accomplished by proper placement of its
poles and zeros. Many well-known techniques for analog and d ­ igital
­filter design can be applied to the synthesis of microring filters.
Thus, a familiarity with classical filter theory is essential for design-
ing high-order microring filters. In this ­section, we will give a brief
review of analog and digital filters and the general ­characteristics
of their transfer functions. The remaining sections of the chapter
100 Optical Microring Resonators

will be devoted to the analysis and synthesis of important classes


of microring filter architectures.
In general filters can be classified into two types: finite impulse
response (FIR) and infinite impulse response (IIR) filters. As the
names suggest, FIR filters are those whose impulse response func-
tion h(t) lasts for a finite duration of time, whereas the impulse
response of an IIR filter continues indefinitely. In integrated optics,
FIR filters are commonly realized using directional couplers and
Mach–Zehnder Interferometers (Madsen and Zhao 1999, Jinguji
and Oguma 2000). IIR filters, on the other hand, are characterized
by some feedback mechanisms and typically require a grating or a
resonator. All microring filters are IIR filters.
The temporal response of an FIR or IIR filter may be continuous
or discrete. A filter whose impulse response is a continuous function
of time is called an analog filter. Its transfer function is given by the
Laplace transform of the impulse response h(t) and is expressed as a
function of the complex frequency variable s. A ­filter whose response
is a discrete function of time is called a digital ­filter. The impulse
response of a digital filter is expressed as a time sequence h(nT), where
n = 0, 1, 2, …, and T is a constant time interval or delay. The z-trans-
form of the impulse response gives the transfer function of the digital
filter, which is expressed as a function of the unit delay variable z.
In Chapter 2 we saw that the transfer function of a microring
resonator obtained using the energy coupling formalism has the
form of a transfer function of an analog filter in the s-domain. On
the other hand, power coupling analysis of the device leads to a
transfer function in the z-domain, where the unit delay variable z−1
represents the round-trip phase delay of the microring. In general,
in the design of microring filters, it is more convenient to synthe-
size s-domain transfer functions using techniques based on the
energy coupling formalism, whereas synthesis of z-domain trans-
fer functions is more naturally performed in the framework of
power coupling. Filter design methods based on energy coupling
are also generally simpler than those based on power coupling.
Another advantage of designing microring filters in the s-domain
is that the transfer functions of a large class of well-known filters
such as Butterworth, Chebyshev, elliptic, and linear phase filters are
readily available. However, since the energy coupling formalism
assumes weak coupling between the microrings, s-domain synthe-
sis methods are strictly accurate only for narrowband filters whose
passbands are much smaller than the FSR of the resonators. For
Coupled Microring Optical Filters 101

broadband or strongly coupled microring filters, a more accurate


design must be sought in the z-domain using the power coupling
approach.
The transfer function describing the transmission response of
an analog IIR filter can be expressed as a rational function in the
s-domain,

∏ (s − z ) ,
M

P(s) K 0 k =1
k
H(s) = = (3.24)
∏ (s − p )
N
Q(s)
k
k =1

where P(s) and Q(s) are polynomials of s, K0 is a constant, and zk and


pk are the zeros and poles, respectively, of the transfer function. The
number of poles, N, gives the order of the filter. The number of zeros,
M, is typically less than or at most equal to N. For ­passive ­filters,
causality restricts the locations of the poles to the left half of the
s-plane. Associated with the transfer function H(s) is a complemen-
tary transfer function F(s), which describes the reflection response
of the filter. The reflection response F(s) is related to the transmis-
sion response H(s) via the Feldtkeller relation for a lossless network,

H(s)H ∗ (− s) + F(s)F ∗ (− s) = 1. (3.25)

Expressing F(s) as a rational function of the form F(s) = R(s)/Q(s),


where R(s) is a polynomial of at most degree N, we obtain from
Equation 3.25

R(s)R∗ (− s) = Q(s)Q∗ (− s) − P(s)P∗ (− s). (3.26)

By factoring the polynomial on the right-hand side of the above


equation, we obtain two sets of roots, one lying in the left half of
the s-plane and the other in the right half. Any combination of N
roots taken from both sets can be used to form the polynomial R(s),
although it is common to choose the roots in the left half-plane. This
choice of zeros yields what is known as the minimum phase reflec-
tion response.
In the inverse z-domain,* the transfer function of a digital IIR
filter has the general form

* For convenience we will work in the inverse z-domain rather the z-domain in
this book. Thus, for example, an Nth-degree polynomial A(z−1) has the general
form A( z −1 ) = a0 + a1z −1 + … + aN − 1z −( N − 1) + aN z − N .
102 Optical Microring Resonators

∏ (z − z ) ,
M
−1
P( z ) −1 K0 k
−1 k =1
H(z ) = = (3.27)
Q( z −1 )
∏ (z − p )
N
−1
k
k =1

where P(z−1) and Q(z−1) are polynomials of z−1, and zk and pk are the
zeros and poles, respectively, in the inverse z-plane. For passive
­filters, the poles pk are restricted to the region outside the unit circle
in the inverse z-plane. The transmission response H(z−1) and the
complementary reflection response, F(z−1) = R(z−1)/Q(z−1), also satisfy
the Feldtkeller relation

H( z −1 )H ∗ ( z) + F( z −1 )F ∗ ( z) = 1, (3.28)

where H*(z) and F*(z) are the para-conjugates of the transfer func-
tions. In terms of the polynomials P, Q, and R, the para-conjugate
transfer functions can be expressed as H ∗ ( z) = P ( z −1 )/Q ( z −1 ) and
F ∗ ( z) = R ( z −1 )/Q ( z −1 ), where P ( z −1 ), Q ( z −1 ), and R ( z −1 ) are the
Hermitian conjugates* of the respective polynomials. Substituting
these expressions into Equation 3.28, we obtain

R( z −1 )R ( z −1 ) = Q( z −1 )Q ( z −1 ) − P( z −1 )P ( z −1 ). (3.29)

The polynomial on the right-hand side of the above equation has two
sets of roots, one lying inside the unit circle and the other lying out-
side. The polynomial R(z−1) can be constructed from any combina-
tion of N of these roots. The choice of the roots lying outside the unit
circle in the inverse z-plane gives the minimum phase filter design.
The design of a microring optical filter begins with the specifi-
cation of the desired transfer functions H and F in either the s- or
inverse z-domain.† The locations of the poles and zeros are chosen

* The Hermitian conjugate (or flip conjugate) of a polynomial A(z−1) of degree N is


defined as A  ( z −1 ) = z − N A∗ ( z), where A*(z) is the para-conjugate (or para-­Hermitian
conjugate) of A(z−1). The Hermitian conjugate can be obtained by reversing
the coefficients of A(z−1) and taking their complex conjugates. For example, if
A( z −1 ) = a0 + a1z −1 + … + aN − 1z −( N − 1) + aN z − N , then its Hermitian conjugate is
A ( z −1 ) = a∗N + a∗N − 1z −1 + … + a1∗ z −( N − 1) + a0∗ z − N
.
† Typically one specifies the desired transmission characteristics of a filter, from
which a suitable transmission transfer function H(s) or H(z−1) is constructed.
Knowing H(s) or H(z−1), one can determine the complementary transfer function
F(s) or F(z−1) for the reflection response using the Feldtkeller relation in Equation
(3.26) or (3.29).
Coupled Microring Optical Filters 103

to achieve specific amplitude, phase, or group-delay characteristics


of the filter. Given a set of prescribed spectral characteristics, the
problem of determining a suitable transfer function is known as the
filter approximation problem, for which a large body of l­iterature
exists in the field of analog and digital filter design (e.g., Lam 1979,
Antoniou 1993, Ellis 1994). From the target transfer functions H and
F, a suitable microring configuration and its coupling coefficients
are then determined, which can reproduce the desired spectral
response. This is called the filter synthesis problem. The remain-
ing sections of this chapter are devoted to developing general
­techniques for analyzing and designing some of the most common
microring filter architectures.

3.3 Cascaded All-Pass Microring Filters


Arrays of cascaded APMRs are the simplest microring architec-
tures for realizing high-order all-pass optical filters. The device
consists of N microring resonators side coupled to the same bus
waveguide as shown in Figure 3.8. A light signal si applied to the
input port couples sequentially to all the microrings. The transmit-
ted ­signal st is the result of the interference between the signals
from all the microrings with the input signal. If there is no intrinsic
loss in the resonators, all the input power will appear at the output
port. Signals at all frequencies will thus pass through the structure
­unattenuated but can acquire a complex phase response. Cascaded
all-pass microring arrays have important applications as phase
­filters, dispersion compensators, and optical delay lines.
To determine the general transfer function of a cascaded all-
pass microring filter, we consider the array of N APMRs shown in
Figure 3.8. Each microring k has radius Rk and is coupled to the
common bus waveguide via field coupling coefficient κk (transmis-
sion coefficient τk). For simplicity we assume that all the microrings
have the same round-trip amplitude attenuation art. We denote the
round-trip phase of microring k as ϕk = ϕ0 + Δϕk, where ϕ0 is a refer-
ence round-trip phase (at a center wavelength λ0) and Δϕk is the phase

R1 R2 R3 RN
κ1 κ2 κ3 κN
si st

Figure 3.8 Schematic of an array of N cascaded APMRs.


104 Optical Microring Resonators

detune of microring k with respect to ϕ0. The transfer function of the


array is simply the product of the transfer functions of individual
APMRs. Using the transfer function of an APMR in Equation 2.44,
we can write the transfer function of the APMR array as
N


s
−1 τ k − art exp(− j∆φ k )z −1
H(z ) = t = , (3.30)
si 1 − τ k art exp(− j∆φ k )z −1
k =1

where z −1 = e − jφ0. Equation 3.30 has the form of the transfer function
of an Nth-order all-pass filter with poles pk = (1/τkart)exp(jΔϕk) and
zeros zk = (τk/art)exp(jΔϕk). Each microring in the array is responsible
for generating a pole and a zero whose locations are determined
by the transmission coefficient τk and the microring phase detune
Δϕk. In the absence of loss, the poles and zeros are related simply by
zk = 1/pk∗.
The phase response ψ of the filter is the sum of the phases of the
N APMRs. Using Equation 2.59 for the phase response ψk of all-pass
microring k, we get
N N N
 art sin φ k   τ k art sin φ k 
ψ= ∑
k =1
ψk = ∑
k =1
tan −1 
 τ k − art cos φ k 
− ∑ tan
k =1
−1
 1 − τ a cos φ  .
k rt k

(3.31)
Similarly, the group-delay response of the array can be obtained
by summing up the group delays of the individual APMRs. Using
Equation 2.62 for the group delay of an APMR, we obtain the group-
delay response of the array as follows:
N N
Sp , k Trt , k Sz , k Trt , k
τg = ∑
k =1
1 + Fk sin 2
( φ k /2)
− ∑ 1 + G sin (φ /2).
k =1 k
2
k
(3.32)

In the above equation, Trt,k is the round-trip delay time of microring


k, and Sp,k and Sz,k are the peak group delays due to the poles and
zeros, respectively,

1  1 + τ k art 
Sp , k =  , (3.33)
2  1 − τ k art 

1  τ k + art 
Sz , k = , (3.34)
2  τ k − art 
Coupled Microring Optical Filters 105

and Fk and Gk are, respectively, given by

Fk = 4τ k art /(1 − τ k art )2 , (3.35)

Gk = 4τ k art /(τ k − art )2 . (3.36)

If all N microrings are identical with identical coupling coefficients,


the phase and group-delay responses of the array are simply N
times those of a single APMR. Such an array can be used as an opti-
cal delay element whose group delay is enhanced by N times over
that of a single APMR.
The cascaded APMR array can be used to synthesize a given
all-pass transfer function of order N by selecting the coupling coef-
ficients and phase shifts of the microrings to realize the required
poles and zeros. As an example, we consider the design of an all-
pass Bessel filter which has a maximally flat group delay response.
In the inverse z-domain, the transfer function of an all-pass Bessel
filter of order N is given by (Prabhu and Van 2008)

∏ (1 − z e ) ,
N
−1 − rk Trt
−1
)=
k =1 (3.37)
∏ (z − e )
H ap ( z N −1 − rk Trt
k =1

where rk are the roots of the Bessel polynomial of degree N. The poles
and zeros of the filter are located at pk = exp(−rkTrt) and zk = exp(rkTrt),
respectively. Assuming that the microrings have very low loss
(art ≈ 1), the transmission coefficient and phase detune of microring
k can be determined from τk = 1/|pk| and ∆φ k = ∠pk, respectively.
For a fifth-order Bessel filter with a 50 GHz bandwidth, the poles of
the filter are located at pk = {1.1045 ± j0.1995, 1.1780 ± j0.1036, 1.2006}.
Assuming that the microrings have an FSR of 1 THz, we compute the
transmission coefficients and phase detunes to be τk = {0.891, 0.891,
0.846, 0.846, 0.833} and Δϕk = {0.179, −0.179, 0.088, −0.088, 0}. Figure 3.9
shows the power transmission and group-delay responses of the
filter in the presence of 1% power loss in the resonators. We see
that the group-delay response has a constant value of 40 ps over
the 50 GHz filter bandwidth. The transmission response is also flat
over this bandwidth, with an insertion loss of −1.75 dB due to loss
in the microrings.
Another important application of cascaded APMR arrays is the
realization of dispersion compensators (Madsen and Lenz 1998).
106 Optical Microring Resonators
(a) (b)
0
40
–1

Group delay, τg (ps)


Transmission (dB)

30
–2

20
–3

–4 10

–5 0
–200 –100 0 100 200 –150 –100 –50 0 50 100 150
Frequency detune, ∆f (GHz) Frequency detune, ∆f (GHz)

Figure 3.9 Spectral responses of (a) power transmission and (b) group
delay of a fifth-order all-pass microring Bessel filter with maximally flat
group delay.

These devices are typically designed to provide a constant chromatic


dispersion D over a bandwidth B. To achieve a group-delay response
with linear slope D, we can use a series of N all-pass microrings
with resonances distributed over the bandwidth B. By designing the
microrings to have increasing or decreasing peak group delays, posi-
tive or negative linear dispersion slopes can be obtained. Figure 3.10
shows an example of a dispersion compensator consisting of five cas-
caded APMRs. The phase detunes of the microrings are fixed at the
values Δϕk = {−0.45π, −0.2π, 0, 0.15π, 0.25π} and the transmission coef-
ficients are set to be τk = {0.333, 0.429, 0.500, 0.556, 0.600}, which are

(a) (b)
1
Normalized group delay, τg/Trt

10
0.8
8
Transmission

0.6
6 Slope D
0.4 1
4 2
3
4
0.2 2 5

0 0
–1 –0.5 0 0.5 1 –1 –0.5 0 0.5 1
Frequency detune, ∆f/FSR Frequency detune, ∆f/FSR

Figure 3.10 Spectral responses of (a) power transmission and (b) normal-
ized group delay (τg/Trt) of a dispersion compensator consisting of five
cascaded APMRs. The curves labeled from 1 to 5 are the group delay
responses of individual microring resonators.
Coupled Microring Optical Filters 107

numerically optimized to give a linear dispersion slope in the group


delay. The microring resonators are assumed to have a round-trip loss
of 2%, which causes a small ripple of about 0.7 dB in the transmission
response of the array (Figure 3.10a). We also see in Figure 3.10b that the
group-delay response exhibits a linear dispersion slope of D = 18Trt/
ΔλFSR over a bandwidth close to half a microring FSR. For example, if
the microrings are designed to have an FSR of 25 GHz (ΔλFSR = 0.2 nm
and round-trip time Trt = 40 ps), the chromatic dispersion D of the
array would be 3.6 ns/nm. Larger values of the dispersion slope can
be obtained by using more microring resonators. The slope D can also
be made negative by simply reversing the sign of each phase detune
value Dfk. A tunable dispersion compensator based on a similar cas-
caded APMR array has been experimentally demonstrated using
Ge-doped silica microrings in Madsen et al. (1999).

3.4 Serially Coupled Microring Filters


Serially coupled microring filters (or CROW filters) are the most
common type of high-order microring filters since these devices are
relatively simple to design and fabricate. Microring filters of vari-
ous orders have been demonstrated in a variety of material systems
including GaAs/AlGaAs (Hryniewicz et al. 2000), SiON (Little et al.
2004), SiN (Barwicz et al. 2004), and SOI (Xia et al. 2007a). In this sec-
tion, we will first develop the theory and techniques for analyzing
and designing serially coupled microring filters using the energy
coupling formalism. This will be followed by a formulation of the
device transfer functions in terms of power coupling, which is more
accurate for broadband or strongly coupled microring devices. A
method for synthesizing serially coupled microring filters using the
power coupling formalism will also be developed.

3.4.1 Energy coupling analysis of serially


coupled microring filters
We consider an array of N serially coupled microring resonators as
shown in Figure 3.11. Each microring i has resonant frequency ωi,
which may be slightly detuned from a center frequency ω0 (typi-
cally taken to be the center frequency of the filter passband). If the
microrings are of approximately the same size, we may assume that
they have the same intrinsic loss. We denote the wave amplitude in
microring i as ai, which is normalized with respect to energy so that
|ai|2 gives the energy stored in the microring. Coupling between
108 Optical Microring Resonators
Input

µ0 µ1 µi µN–1 µN
si
ai
1 2 i i+1 N–1 N
st sd

Through Drop

Figure 3.11 Schematic of a serially coupled microring filter consisting of


N microring resonators.

two adjacent microrings i and i + 1 is denoted by the ring-to-ring


energy coupling coefficient μi (1 ≤ i ≤ N−1). In addition, microrings 1
and N are also coupled to an input waveguide and an output wave-
guide via ring-to-bus coupling coefficients μ0 and μN, respectively.
An input signal si carrying power |si|2 is applied to the input port
2
which supplies energy to microring 1 at a rate of µ 02 si . At the same
time, energies in microrings 1 and N are coupled out to the input
and output bus ­waveguides, respectively, and transmitted to the
through port and drop port as power-normalized waves st and sd.
In each microring resonator the rate of change of energy is due
to a combination of intrinsic loss and coupling to neighbor micror-
ings, as well as coupling to the bus waveguides in the case of
microrings 1 and N. We can thus write the coupled mode equations
describing energy transfer between the microrings as (Haus 1984,
Little et al. 1997)
da1
= ( jω1 − γ i − γ 0 )a1 − jµ1 a2 − jµ 0 si ,
dt
da2
= ( jω 2 − γ i )a2 − jµ1 a1 − jµ 2 a3 ,
dt
… (3.38)
daN −1
= ( jω N −1 − γ i )aN −1 − jµ N − 2 aN − 2 − jµ N −1 aN ,
dt
daN
= ( jω N − γ i − γ N )a − jµ N −1 aN −1 .
dt

In the above equations, γi is the amplitude decay rate due to intrin-


sic loss, which is related to the propagation loss α in the micror-
ing waveguide by γi = αvg/2, where vg is the group velocity. The
decay rates γ0 and γN are due to coupling from microrings 1 and N
Coupled Microring Optical Filters 109

to the input and output bus waveguides, respectively, and can be


computed from the ring-to-bus energy coupling coefficients using
Equation 2.73 as γ 0 = µ 02 /2 and γ N = µ 2N /2.
In microring filter design, it is necessary to relate the energy
coupling coefficients to the field coupling coefficients, which are
then used to determine the coupling gaps between the microrings
or between the microrings and the bus waveguides. The ring-to-
bus energy coupling coefficients can be computed from the field
coupling coefficients κ0 and κN using Equation 2.76,

vg κi
µi = κ i = , i = 0, N , (3.39)
2πRi Trt , i

where Ri is the radius of microring i and Trt,i is its round-trip time.


To relate the ring-to-ring energy coupling coefficients μi (1 ≤ i ≤ N−1)
to the corresponding field coupling coefficients κi, we observe that
the rate of energy coupling from ring i+1 to ring i is given by

2  2πRi +1  2 2
µ i2 ai +1 (t) = µ i2  2
 Ai +1 (t) ≡ g i Ai +1 (t) , (3.40)
 vg 
2
where |Ai+1(t)|2 is the power circulating in microring i+1, and g i2 Ai +1
is the rate at which the power-normalized wave Ai+1 supplies energy
to microring i. We can regard the wave Ai+1 as playing the same role
as the power-normalized wave si, which supplies energy to microring
2
1 at a rate equal to µ 02 si . Using Equation 3.39, we can compute gi from
the field coupling coefficient κi as g i = κ i vg /2πRi . We thus have

g i2 = µ i2 (2πRi +1 /vg ) = κ i2 (vg /2πRi ),

which yields

vg κi
µi = κ i = , 1 ≤ i ≤ N − 1. (3.41)
2π Ri Ri +1 Trt ,i Trt ,i +1

If the microrings are identical with the same round-trip time, the
above equation simplifies to μi = κi/Trt.
To obtain the transfer functions of the serially coupled micror-
ing filter, we solve the coupled mode equations in Equation 3.38 for
the case of a harmonic input wave excitation of the form si ~ ejωt. In
this case the solutions for the wave amplitudes ai in the microrings
110 Optical Microring Resonators

will also vary as ai ~ ejωt. Applying these solutions to Equation 3.38,


we get
[ j(ω − ω1 ) + γ i + γ 0 ]a1 + jµ1a2 = − jµ 0 si ,
jµ1a1 + [ j(ω − ω 2 ) + γ i ]a2 + jµ 2 a3 = 0,
… (3.42)
jµ N −2 aN −2 + [ j(ω − ω N −1 ) + γ i ]aN −1 + jµ N −1aN = 0,
jµ N −1aN −1 + [ j(ω − ω N ) + γ i + γ N ]aN = 0.

It is convenient to write the terms ω − ωi in the above equations as

ω − ω i = (ω − ω 0 ) − (ω i − ω 0 ) = ∆ω − δω i , (3.43)

where Δω is the frequency detune from the center frequency ω0 and


δωi is the deviation of the resonance of microring i from ω0. Defining
the complex frequency s = jΔω + γi, we can express Equation 3.42 in
the matrix form
Ka = s, (3.44)
where a = [a1, a2, …aN−1, aN]T is the wave amplitude array, s = [−jμ0 si,
0, …0, 0]T is the input excitation vector, and K is a symmetric tridi-
agonal matrix,

 s − jδω1 + γ 0 jµ1 
 jµ1 s − jδω 2 jµ 2 
 
K= ⋅ ⋅ ⋅ .
 
 jµ N − 2 s − jδω N −1 jµ N −1 
 jµ N −1 s − jδω N + γ N 
(3.45)
For a given input wave amplitude si, the solution of the matrix
Equation 3.44 gives the wave amplitudes ai in the microrings.
At the drop port of the microring filter, the transmitted signal is
related to the wave amplitude in microring N by sd = −jμNaN. We can
thus write the drop port transfer function of the filter as

sd − jµ N aN
Hd = = . (3.46)
si si

From Equation 3.44, we obtain the solution for aN as aN = det(A)/


det(K), where A is the matrix obtained by replacing the last column
Coupled Microring Optical Filters 111

of K with s. The determinant of A can be explicitly computed as


follows:
N −1
det(A) = − jµ 0 si ∏ (− jµ k ) = (− j)N (µ 0 µ1µ 2  µ N −1 )si . (3.47)
k =1

For the tridiagonal matrix K, its determinant is given by det(K) = CN,


where CN is the Nth-continuant obtained from the recursive formula*

* Ck = K N − K +1,N − K +1Ck −1 + µ 2N − K +1Ck − 2 , (k ≥ 2) (3.48)

with C0 = 1 and C1 = K N,N. The drop port transfer function of the


microring filter can thus be expressed as

(− j)N +1 (µ 0 µ1µ 2  µ N )
Hd (s) = , (3.49)
CN (s)

where the continuant CN(s) has the form of a polynomial of degree N.


To determine the transfer function at the through port of the fil-
ter, we observe that the through port signal is given by st = si − jμ0 a1.
We can thus write the through port transfer function as
st jµ 0 a1
Ht = = 1− . (3.50)
si si
The wave amplitude in microring 1 can be obtained from a1 = det(B)/
det(K), where B is the matrix obtained by replacing the first column of
K with s. The solution can be expressed in terms of the continuants as
CN −1 (s)
a1 = − jµ 0 si , (3.51)
CN (s)

which may also be written in the form of a continued fraction,


− jµ 0 si
a1 = .
µ12
s − jδω1 + γ 0 +
µ 22 (3.52)
s − jδω 2 +
µ 2N −1
s − jδω 3 + …
s − jδω N + γ N

* Equation 3.48 expresses the continuant for the matrix K whose columns and
rows have been flipped left-to-right and up-to-down, respectively. This allows
the same recursive formula to be used to calculate the determinant of the matrix
B in Equation 3.51.
112 Optical Microring Resonators

The above solution for a1 can also be obtained through a back-


ward substitution procedure of the system Ka = s. Using the result
in Equation 3.52, we obtain the transfer function at the through
port as

µ 02
H t (s) = 1 − .
µ12
s − jδω1 + γ 0 +
µ 22
s − jδω 2 +
µ 2N −1
s − jδω 3 + …
s − jδω N + γ N
(3.53)
Equations 3.49 and 3.53 give the closed-form expressions for the
transfer functions at the drop port and through port of an Nth-order
serially coupled microring filter. We note that the drop port transfer
function has N poles but no finite zeros. Thus, the serial microring
coupling configuration can only be used to realize all-poles transfer
functions, such as those of Butterworth (or maximally flat) filters,
Chebyshev filters, and Bessel filters.
If the output bus waveguide is removed from the microring
array in Figure 3.11, the structure becomes an Nth-order all-pass
microring filter. The transfer function Ht = st/si of the device is given
by Equation 3.53 with the output bus-to-ring coupling coefficient
μN set to 0 (so that γN = 0). In the all-pass configuration, the seri-
ally coupled microring filter is equivalent to the cascaded all-pass
microring array in Section 3.3 and can be used to realize all-pass
transfer functions of any order N.

3.4.2 Energy coupling synthesis of serially


coupled microring filters
In this section we develop a simple method for synthesizing seri-
ally coupled microring filters to achieve a target spectral response.
The method exploits the fact that the through port transfer function
Ht(s) of the microring filter can be expressed as a continued fraction
(Prabhu et al. 2008a). Suppose that the target filter has an Nth-order
through port (or reflection) transfer function given by

R(s) s N + rN −1s N −1 +  + r1s + r0


H t (s) = = . (3.54)
Q(s) s N + qN −1s N −1 +  + q1s + q0
Coupled Microring Optical Filters 113

We form the expression

Q(s) − R(s) M (s)


1 − H t (s) = = (qN −1 − rN −1 ) N −1 , (3.55)
Q(s) Q(s)

where MN−1(s) is a polynomial of degree N−1 with the leading


­coefficient equal to 1. Assuming that the microrings are synchro-
nously tuned (δωi = 0) and lossless, we rearrange Equation 3.53
to read

µ 02
1 − H t (s) = ,
µ12
(s + γ 0 ) +
µ 22 (3.56)
s+
µ2
s + … N −1
s+ γN

where s = jΔω. Comparing Equations 3.55 and 3.56 and noting that
the leading coefficients of both MN−1(s) and Q(s) are 1, we obtain
µ 02 = qN −1 − rN −1. By expressing MN−1(s)/Q(s) in the form of a contin-
ued fraction, we can determine the remaining coupling coefficients
μk. Specifically, carrying out the division

Q(s) b ⋅ MN − 2 (s)
= (s + γ 0 ) + (3.57)
M N −1 (s) MN −1 (s)

where γ 0 = µ 02 / 2 and MN−2(s) are polynomial of degree N–1, we


obtain b = µ12. Next, by dividing MN−1(s) by MN−2(s), we can get μ2.
This process is repeated until all the coupling coefficients are deter-
mined. This method can also be used to synthesize all-pass micror-
ing filters consisting of N serially coupled microring resonators.
Two common types of filters that can be realized with the serial
coupling configuration are the Butterworth and Chebyshev filters.
For these filters, closed-form design formulas for computing the
energy coupling coefficients can be obtained. Butterworth filters
(also called maximally flat filters) of order N have the property that
the first (2N–1) derivatives of the frequency response |H(jω)|2 are
zero at the center frequency. Chebyshev filters provide steeper skirt
roll-off than Butterworth filters but have equi-ripples in the pass-
band. The ripples are characterized by a small positive parameter
ε such that |H(jω)|2 oscillates between 1 and 1/(1 + ε2) within the
114 Optical Microring Resonators

passband. The transfer function of an Nth-order Butterworth or


Chebyshev filter can be expressed as
N
1 1
H(s) =
K0 ∏s− p ,
k =1 k
(3.58)

where s = 2j(ω − ω0)/ΔωB and ΔωB is the filter bandwidth. For


Butterworth filters, the constant K0 is equal to 1 and ΔωB is the 3 dB
bandwidth. For Chebyshev filters, K0 = 2N−1ε and ΔωB is the ripple
bandwidth. The poles pk of the filter are determined by
pk = − a sin θ k + jb cos θ k ,
(2k − 1)π (3.59)
θk = .
2N

For Butterworth filters, a = 1 and b = 1, whereas for Chebyshev fil-


ters they are given by

1  1 
a = sinh  sinh −1    ,
N  ε 
(3.60)
1  1 
b = cosh  sinh −1    .
N  ε 

Butterworth and Chebyshev filters of order N can be realized


using N serially coupled microring resonators. The energy coupling
coefficients can be computed directly from the formulas (Van 2006)

c0 ∆ω B /2
µ 02 = µ 2N = ,
sin( π/2N )
(3.61)
(ck ∆ω B /4)2
µ k2 = , 1 ≤ k ≤ N −1
sin [(2k − 1)π/2N ] sin [(22k + 1/2N )π ]

In the above expressions, ck = 1 for Butterworth filters and

 kπ   1  1 + ε 2 + 1 
ck2 = sin 2   + sinh 2  ln  , 0 ≤ k ≤ N (3.62)
 N  2N  1 + ε 2 − 1  

for Chebyshev filters with ripple parameter ε.


As an example, we consider the design of a fourth-order
Butterworth filter and a Chebyshev filter of the same order with a
Coupled Microring Optical Filters 115

0.5 dB ripple. The filter bandwidth is specified to be Δf B = 50 GHz.


Using Equation 3.61, we compute the energy coupling coefficients to
get μ0 = μ4 = 20.26 GHz1/2, {μ1, μ2, μ3} = {132.09, 85.01, 132.09} GHz for
the Butterworth filter, and μ0 = μ4 = 13.71 GHz1/2, {μ1, μ2, μ3} = {111.30,
93.51, 111.30} GHz for the Chebyshev filter. Using microrings with
FSR = 1 THz, we calculate the corresponding field coupling coef-
ficients using Equations 3.39 and 3.41 to be {κ0, κ1, κ2, κ3, κ4} = {0.641,
0.132, 0.085, 0.132, 0.641} for the Butterworth filter and {κ0, κ1, κ2,
κ3, κ4} = {0.434, 0.111, 0.094, 0.111, 0.434} for the Chebyshev filter. In
Figure 3.12 we plot the spectral responses at the drop port and
through port of both filters. The Chebyshev filter is seen to have a
sharper roll-off than the Butterworth filter at the expense of a small
ripple in the passband. In general, the roll-off can be made steeper
at the expense of increased in-band ripples.
Due to their flat-top passbands and relatively steep roll-offs,
high-order Butterworth microring filters are especially of interest
for applications as add-drop filters in WDM communication sys-
tems. Filters of various orders have been fabricated in a variety
of material systems. One of the earliest experimental results that
demonstrated the viability of these devices is shown in Figure 3.13.
The plot shows the measured and theoretical spectral responses of
Butterworth microring filters of increasing orders from 1 to 6 fabri-
cated in the Hydex material system (Van et al. 2004). The microrings
had a radius of 42 µm and the filters were designed to have a 3 dB
bandwidth of around 100 GHz. Fairly good agreement between the
measured and designed responses can be seen. It is also evident

0
Transmission (dB)

–10

–20
Drop

–30

–40 Through

–50
–100 –50 0 50 100
Frequency detune, ∆f (GHz)

Figure 3.12 Spectral responses of fourth-order serially coupled


microring filters: Butterworth filters (black lines) and Chebyshev filters
(gray lines).
116 Optical Microring Resonators

0 Dashed—measured
Solid—ideal max. flat

–10
N=1

Transmission (dB)
–20
N=2
–30

–40 N=3

N=5 N=4
–50
N=6

–60
–3 –2 –1 0 1 2 3
Normalized frequency, ∆f/∆f3dB

Figure 3.13 Measured and theoretical spectral responses of Butterworth


microring filters of orders 1–6 fabricated in the Hydex material.
(From Van, V., et al., 2004. Micro-ring resonator filters. In The 17th Annual
Meeting of the IEEE Lasers and Electro-Optics Society, Vol. 2, pp. 571–572.)

that the filter passband becomes flatter and the roll-off becomes
steeper as the filter order is increased.

3.4.3 Power coupling analysis of serially


coupled microring filters
The energy coupling methods developed in Sections 3.4.1 and
3.4.2 are strictly valid only for narrowband microring filters,
that is, filters whose 3 dB bandwidths are much smaller than the
FSRs of the microrings. For broadband or, equivalently, strongly
coupled microring filters, a more accurate analysis or design
must be performed using the power coupling formalism. In this
approach, the microring device is treated as a digital IIR filter,
whose ­t ransfer functions are specified in terms of the delay vari-
able z−1.
In this section, we apply the transfer matrix method developed
in Section 3.1.3 for infinite CROW arrays to obtain the z-domain
transfer functions of a serially coupled microring filter of order N
(Orta et al. 1995, Poon et al. 2003). For simplicity, we assume that
the microrings are identical with the same radius R, propagation
­constant β, and loss α in the microring waveguides. Coupling
between adjacent microrings i and i+1 is denoted by the field
­coupling coefficient κi (1 ≤ i ≤ N−1). Microrings 1 and N are also
Coupled Microring Optical Filters 117

coupled to the input and output bus waveguides via field coupling
coefficients κ0 and κN, respectively.
In each microring i, we label the fields ai, bi, ci, di in the direction
of wave propagation, as shown in Figure 3.14. The four fields are
related by a transfer matrix P defined as

ci   0 art1/2 e − jφrt /2  ai  1/2 0 z −1  ai  1/2  i 


a
=
d   −1/2 jφrt /2   = z    ≡ z P ,
 i  art e 0  bi  1 0  bi  bi 
(3.63)

where z −1 = art e − jφrt, ϕrt = 2πβR is the round-trip phase and art = e−παR
is the round-trip attenuation factor in the microrings. The coupling
junction between microrings i and i+1 is described by the transfer
matrix Ki given in Equation 3.21,

 ai +1  1  τ i −1   ci   ci 
 b  = jκ  1   
− τ i   di 
≡ Ki   , (3.64)
 i +1  i   di 

where τ i = 1 − κ i2 . Combining Equations 3.63 and 3.64, we obtain


the transfer matrix for each unit element i in the microring array as
follows:

 ai +1  1/2  ai 
b  = z K iP b  . (3.65)
 i +1   i

The transfer matrix T of the array of N microrings is then

 sa  N /2  si  N /2  i 
s
 s  = z (K N P)(K N −1P)(K1P)K 0  s  ≡ z T  s  . (3.66)
 d  t  t

Element i
κ0 κN
si κi
a1 d1 ai+1 di+1 cN sa
bi ci

st b1 c1 ai di bi+1 ci+1 dN sd

1 i i+1 N

Figure 3.14 Schematic of N serially coupled microrings for transfer


matrix analysis.
118 Optical Microring Resonators

The transfer functions at the drop port and through port of the
microring filter are obtained by setting sa = 0 in Equation 3.66 and
solving for st/si and sd/si. The results are

st T
H t ( z) = = − 11 , (3.67)
si T12

sd  T T  det(T )
Hd ( z) = = z N/2  T21 − 11 22  = − z N /2 . (3.68)
si  T12  T12

The determinant of T is just the product of the determinants of the


matrices Ki and P in Equations 3.63 and 3.64. Since det(P) = −z−1 and
det(Ki) = −1, we obtain det(T) = −z−N. With this result, we can write
the transfer function at the drop port as

z − N /2
Hd ( z) = . (3.69)
T12

If we define the transfer matrix Tk of the first k microrings as

Tk = (K k P)(K k −1P)(K1P)K 0 , (3.70)

then we have the recursive formula, Tk = (K kP)Tk−1. Starting from


T0 = K0 and applying the recursive formula, we find that the trans-
fer matrix T = TN of N serially coupled microrings has the general
form

1  RN ( z −1 ) σ N QN ( z −1 )
TN =  , (3.71)
( jκ 0 )( jκ 1 )( jκ N ) Q N ( z −1 ) σ N R N ( z −1 ) 

where R N and QN are polynomials of degree N, R N , and Q N are


their Hermitian conjugates and σN = +/−1 if N is odd/even. The
polynomial QN has the leading coefficient equal to 1 (coefficient
of the z−N term). The polynomials R N and QN obey the recursive
relations

Rk ( z −1 ) = τ k z −1Q k −1 ( z −1 ) − Rk −1 ( z −1 ), (3.72)

Qk ( z −1 ) = Qk −1 ( z −1 ) − τ k z −1R k −1 ( z −1 ), (3.73)
Coupled Microring Optical Filters 119

for 1 ≤ k ≤ N with R0 = τ0 and Q0 = 1. Using Equations 3.67 and 3.69,


we can write the transfer functions of a serially coupled microring
filter of order N as

( jκ 0 )( jκ 1 )( jκ N )
Hd ( z −1 ) = σ N z − N/2 , (3.74)
QN ( z −1 )

RN ( z −1 )
H t ( z −1 ) = − . (3.75)
QN ( z −1 )

The above equations show that the drop port and through port
transfer functions of the filter have N poles, which are given by
the roots of the polynomial QN. The drop port has no transmission
zeros (other than the trivial zeros at z−1 = 0), while the through port
has N zeros given by the roots of RN.

3.4.4 Power coupling synthesis of serially


coupled microring filters
Given the transfer functions of an all-poles filter in the inverse
z-domain, it is possible to determine the coupling coefficients of
a serially coupled microring filter that will reproduce the desired
spectral responses. The filter synthesis method we develop here is
similar to the technique first introduced in Orta et al. (1995) and
is based on the transfer matrix method and order reduction tech-
nique. In the method, we start from the Nth-order transmission and
reflection transfer functions of the target filter and extract the cou-
pling coefficients of the unit elements (microrings) in the array one
by one starting with the last coefficient (κN). Each time a coupling
coefficient is extracted, the order of the transfer functions is reduced
by one (i.e., the array is reduced by one unit element), and the proce-
dure is repeated until the first coefficient (κ0) is determined.
Suppose we know the polynomials Rk and Qk of the transfer
matrix Tk of the first k microrings. From the recursive relations
(3.72) and (3.73), we solve for Rk−1 and Qk−1 to get

τ k Q k ( z −1 ) − Rk ( z −1 )
Rk −1 ( z −1 ) = , (3.76)
κ 2k

Qk ( z −1 ) − τ k R k ( z −1 )
Qk −1 ( z −1 ) = . (3.77)
κ 2k
120 Optical Microring Resonators

Since Rk and Qk are polynomials of degree k while Rk−1 and Qk−1


have degree k–1, the coefficient of the z−k term on the right-hand
side of Equations 3.76 and 3.77 must be zero. By imposing this con-
dition, we obtain the following expression for the transmission
coefficient τk:

rk( k ) qk( k )
τk = = , 0 ≤ k ≤ N, (3.78)
q0( k ) r0( k )

(k )
where rk( k ) and qk denote the coefficients of the kth-power terms of
Rk and Qk, respectively.
In the design of a serially coupled microring filter, we assume
that the drop port and through port transfer functions of the target
filter are given by Hd(z−1) = K0/Q(z−1) and Ht(z−1) = R(z−1)/Q(z−1), where
K0 is a constant and R(z−1) and Q(z−1) are polynomials of degree N of
the form

R( z −1 ) = rN z − N + rN −1 z −( N −1) + r1 z −1 + r0 , (3.79)

Q( z −1 ) = z − N + qN −1 z −( N −1) + q1 z −1 + q0 . (3.80)

We begin by calculating the transmission coefficient τN = rN/q0 = 1/r0


of the coupling junction between microring N and the output wave-
guide. Knowledge of τN and κN allows us to compute R N−1 and QN−1
using Equations 3.76 and 3.77. The transmission c­ oefficient τN−1 is
next obtained using Equation 3.78 and the p ­ rocedure is repeated
until all the coupling coefficients are obtained.
The above method requires that the transfer functions of the
target filter be specified in terms of the z−1 variable. Two impor-
tant types of all-poles filters whose transfer functions can be read-
ily obtained in the inverse z-domain are the Butterworth and
Chebyshev filters. Specifically, Butterworth filters have all reflection
zeros (roots of R(z−1)) located at zk = 1. On the other hand, the reflec-
tion zeros of a Chebyshev filter of order N are located at (Orta et al.
1995)

   ∆ω R  (2k − 1)π  
zk = − exp 2 j cos −1 sin  cos  , (k = 1… N )


  4∆ω FSR  2N  
(3.81)
Coupled Microring Optical Filters 121

where ΔωR and ΔωFSR are the ripple bandwidth and the FSR, respec-
tively. From the given zeros, we can construct the polynomial R(z−1)
as follows:
N
R( z −1 ) = ∏ (z
k =1
−1
− zk ). (3.82)

Using the Feldtkeller relation in Equation 3.29, we have

Q( z −1 )Q ( z −1 ) = R( z −1 )R ( z −1 ) + z − N K 02 , (3.83)

where the constant K0 can be found from the 3 dB bandwidth


ΔωB of the filter. Specifically, writing zB−1 = exp(− j∆ω B /2∆ω FSR ),
we have

2 K 02 1
Hd ( zB−1 ) = = , (3.84)
2 −1 2 2
K + R( z )
0 B

from which we obtain K 0 =|R( zB−1 )|. To determine Q(z−1), we find the
roots of the polynomial on the right-hand side of Equation 3.83 and
choose N roots with magnitude greater than unity* from which to
construct the polynomial Q(z−1).
As an example, we consider the design of a fifth-order micror-
ing Chebyshev filter with 3 dB bandwidth ΔωB = 0.2ΔωFSR and
ripple bandwidth ΔωR = 0.1ΔωFSR. We use Equation 3.81 to com-
pute the zeros of the polynomial R(z−1) and determine its coeffi-
cients to be {r0, …, r5} = {−2.3874, 11.6448, −23.0047, 23.0047, −11.6448,
2.3874}. From the 3 dB bandwidth, we calculate K0 = 0.1507 using
Equation 3.84. Next we construct the polynomial Q(z−1) using the
Feldtkeller relation in Equation 3.83 and obtain the coefficients
{q0, …, q5} = {−5.6996, 18.2328, −24.8563, 17.7288, −6.5563, 1}. The filter
is realized with an array of five serially coupled microring reso-
nators with coupling coefficients {κ0, …, κ5} = {0.908, 0.522, 0.343,
0.343, 0.522, 0.908}. The target spectral responses and those at the
drop port and through port of the synthesized filter are shown in
Figure 3.15.

* We choose roots lying outside the unit circle since we are operating in the
inverse z-domain.
122 Optical Microring Resonators

3.5 Parallel Cascaded ADMRs


In Section 3.1.2 we showed that an infinite array of cascaded
ADMRs has transmission bands and stopbands that arise from
both the microring and Bragg resonances. For an ADMR array of
finite length, the transmission band has ripples due to impedance
mismatch or reflections at the two terminations. By apodizing the
coupling coefficients of the ADMRs, the ripples can be reduced or
eliminated to produce a filter response with flat-top passband and
smooth skirt roll-offs (Little et al. 2000). Apodization is a technique
commonly employed in the design of Bragg grating filters (Hill
and Meltz 1997). For a microring array, we can apodize either the
input coupling coefficients (i.e., the coupling coefficients between
the microrings and the input waveguide) or the output coupling
coefficients, or both. Apodizing both coupling coefficients results
in symmetrically coupled ADMRs which are simpler to analyze. In
this section we use the transfer matrix method to derive the trans-
fer functions of a finite array of cascaded ADMRs with symmetric
couplings (Grover et al. 2002).
Figure 3.16 shows a schematic of an apodized array of N cascaded
ADMRs. Each microring k in the array is coupled to the input and
output bus waveguides with equal input and output field coupling
coefficients κk. For simplicity, we also assume that all the microrings
are identical with round-trip phase ϕrt and amplitude attenuation art.
Adjacent microrings are separated by two parallel waveguide sec-
tions of length L, which is assumed to be greater than the microring

–10
Transmission (dB)

–20 Drop
|Hd|2 Through
–30 |Ht|2

–40

–50

–60

–1 –0.5 0 0.5 1
Normalize frequency detune, ∆f/FSR

Figure 3.15 Spectral responses of a fifth-order Chebyshev serially cou-


pled microring filter (black solid lines: target filter response; gray dashed
lines: synthesized filter responses).
Coupled Microring Optical Filters 123

diameter so that there is no direct coupling between the two micror-


ings. An input signal si applied to the input waveguide will be
partially coupled to the microrings and the remaining power is
transmitted to the through port. The light waves in the microrings
are in turn coupled to the output waveguide where they combine to
give the drop port signal sd. Note that the signal that emerges at the
drop port of each microring propagates backward to the previous
microring, thereby forming an effective cavity of length L between
two adjacent resonators. This feedback gives rise to additional reso-
nances besides those of the microring resonators and, as a result, the
transfer functions of the ADMR array will contain more poles than
the number of microrings in the array.
With reference to Figure 3.16, we characterize each ADMR k in
the array by a transfer matrix Mk defined as

 ck   ak 
 d  = Mk  b  . (3.85)
 k  k

The elements of Mk are given by

(k ) τ k2 − z −1
M11 = , (3.86)
τ k (1 − z −1 )

(k ) (k ) −κ 2k z −1/2
M21 = − M12 = , (3.87)
τ k (1 − z −1 )

(k ) 1 − τ 2k z −1
M22 = , (3.88)
τ k (1 − z −1 )

si ak ck ak+1 ck+1 st
Input waveguide L
Input Through
1 2 k k+1 N
Drop
Output waveguide L
sd bk dk bk+1 dk+1

Figure 3.16 Schematic of an array of N cascaded add-drop microring


resonators with coupling apodization.
124 Optical Microring Resonators

where z −1 = art e − jφrt and ϕrt = βr2πR is the round-trip phase of the
microring. The transfer matrix P of the two parallel waveguides of
length L connecting two adjacent microrings is

 ak + 1   e 0   ck 
− jθ
 ck 
b   = 
jθ  d 
≡ P , (3.89)
 k + 1   0 e   k   dk 

where θ = βbL is the phase shift and βb is the propagation constant


of the bus waveguides. For simplicity we have neglected loss in the
bus waveguides. The total transfer matrix of the ADMR array is
simply the product of the transfer matrices of the N stages,

T = (M N P)(M N −1P)(M2P)M1 . (3.90)

The transfer functions at the drop port and through port of the
array are then obtained from Hd = −T21/T22 and Ht = det(T)/T22,
respectively.
Analytical expressions for the transfer functions of a cascaded
ADMR array can be obtained for certain values of the spacing L.
For the special case where L = πR + λ0/4nb (Little et al. 2000), we
can write θ ≈ ϕrt/2 + π/2, where we have assumed that the propaga-
tion constants of the bus waveguides and microring waveguides
are equal (βb ≈ βr), and that the phase term βr(λ0/4nb) ≈ π/2 changes
much more slowly over one FSR compared to the round-trip phase
ϕrt of the microring. With these approximations, we have e−jθ = −jz−1/2,
so the matrix P in Equation 3.89 becomes

 − jz −1/2 0   −1
1/2 − z 0
P=  = jz  . (3.91)
 0 jz1/2   0 1

If we define the transfer matrix of the first k microring stages as

Tk = (Mk P)(Mk −1P)(M2P)M1 , (3.92)

then starting from T1 = M1, we can use the recursive relation


Tk = (MkP)Tk−1 to show that the transfer matrix TN = T of the entire
array has the general form

( jz1/2 )N −1  z −1RN ( z −1 ) − z −1/2 PN ( z −1 )


T=  −1/2 , (3.93)
τ1τ 2  τ N (1 − z −1 )N  z PN ( z )
−1
− R N ( z −1 ) 
Coupled Microring Optical Filters 125

and the polynomials in the matrix can be computed from the


­recursive relations

Rk = −(τ 2k − z −1 )z −1Rk −1 − κ k2 Pk −1 , (3.94)

R k = (1 − τ 2k z −1 )R k −1 − κ 2k Pk −1 z −2 , (3.95)

Pk = −κ 2k z −2 Rk −1 + (1 − τ 2k z −1 )Pk −1 , (3.96)

Pk = −κ 2k R k −1 − (τ 2k − z −1 )z −1Pk −1 . (3.97)

The transfer functions at the drop port and through port of the cas-
caded ADMR array are obtained from the matrix T as follows:

T21 z −1/2 P
Hd ( z −1 ) = − =−  N , (3.98)
T22 RN

det(T ) (1 − z −1 )N
H t ( z −1 ) = = τ1τ 2  τ N ( jz)−( N −1)/2 . (3.99)
T22 R N

Equation 3.99 indicates that the through port transfer function has
N identical zeros at z−1 = 1, implying that the drop port response will
have a maximally flat passband. On the other hand, since PN and R N
are polynomials of degree 2(N – 1), the drop port transfer function
in Equation 3.98 has 2(N – 1) zeros and 2(N – 1) poles. Since there
are more poles and zeros in the transfer function than the number
of design parameters (which are the N coupling coefficients), the
poles and transmission zeros of the filter cannot be independently
chosen.
In Figure 3.17a we plot the spectral responses at the drop port
and through port of an unapodized array of N = 7 cascaded ADMRs
with identical coupling coefficients κk = 0.1. The waveguide length
L is chosen to be L = πR + λ0/4nb. We observe that a flat-top pass-
band occurs at the resonant frequency of the microrings, but there
are large side lobes in the drop port response. To suppress the side
lobes, we apply apodization to the coupling coefficients in the array
according to the Gaussian function κk = 0.1exp[−0.25(k − 4)2] (Little
et al. 2000). The spectral responses of the apodized ADMR array are
plotted in Figure 3.17b, which shows that the side lobes have been
suppressed to produce a flat-top filter response with a smooth skirt
roll-off.
126 Optical Microring Resonators
(a) (b)
0 0
|Ht|2 |Ht|2

Transmission (dB)
Transmission (dB)
–10 –10
|Hd|2
–20 –20 |Hd|2

–30 –30

–40 –40

–50 –50
–0.015 –0.01 –0.005 0 0.005 0.01 0.015 –0.015 –0.01 –0.005 0 0.005 0.01 0.015
Roundtrip phase detune, ∆φrt Roundtrip phase detune, ∆φrt

Figure 3.17 (a) Drop port and through port spectral responses of an array
of seven cascaded ADMRs (a) with no coupling apodization (κk = 0.1) and
(b) with Gaussian apodization of the coupling coefficients.

3.6 Parallel Cascaded Microring Doublets


In the previous section we saw that in an array of cascaded ADMRs,
the drop port and through port signals of the microrings propagate
in opposite directions in the two bus waveguides. This forms feed-
back loops in the array which give rise to more poles (or resonances)
in the transfer functions than the number of resonators in the struc-
ture. Since the poles and zeros cannot be independently placed, it
is generally not possible to exactly synthesize a given filter transfer
function using cascaded ADMRs.
Arrays of cascaded ADMRs are examples of a feedback micror-
ing array, which consists of a parallel cascade of microring networks,
typically of odd orders, whose output signals at the drop port and
through port travel in opposite directions in the two connecting bus
waveguides. Feedback microring arrays share the common feature
that their transfer functions have more poles than the number of res-
onators in the structure. On the other hand, a feedforward microring
array is a cascade of microring networks in which the output signals
at the drop port and through port of each microring stage propagate
in the same direction in the two connecting bus waveguides (Prabhu
et al. 2008b). Since there is no feedback in the structure, the transfer
functions of a feedforward array contain exactly the same number
of poles as the number of microring resonators. These structures are
more useful for filter applications than feedback arrays since they
can be used to exactly realize a given transfer function.
The simplest example of a feedforward microring array is a par-
allel cascaded array of symmetric microring doublets, as shown in
Coupled Microring Optical Filters 127

si γ2 ak ck γk + 1 ak + 1 γN st
Input Through
κ1, 1 κ1, k κ1, N
L L
1 2k–1 2N–1

κ2, 1 κ2, k κ2, N

2 2k 2N
κ1, 1 L κ1, k L κ1, N
Drop
bk dk bk + 1 sd

Stage 1 Stage k Stage k + 1 Stage N

Figure 3.18 Schematic of a microring ladder filter consisting of N


­cascaded microring doublets.

Figure 3.18 (Liew and Van 2008). We refer to this structure simply
as a microring ladder filter. As we will show below, the transfer
function of a microring ladder filter contains transmission zeros
that make them useful for realizing new classes of filters, such as
inverse Chebyshev, pseudo-elliptic,* and linear phase filters, that
are not possible with conventional CROW filters.
We will use the power coupling formalism and the transfer
matrix method to analyze the microring ladder filter in Figure 3.18.
The array consists of N microring doublet stages, for a total of 2N
microring resonators. We assume the microrings to be lossless and
synchronously tuned to the same resonant frequency. Each micror-
ing doublet in stage k is composed of two serially coupled micror-
ing resonators with symmetric bus-to-ring coupling coefficients
κ1,k and ring-to-ring coupling coefficient κ2,k. Connecting adjacent
microring stages are two parallel bus waveguides of length L, with
a possible π-phase shift in the upper waveguide with respect to
the lower waveguide. We denote this π-phase shift by the factor
γk = e−jπ = −1. If there is no phase shift, then γk = 1. The differential
π-phase shifts are ­necessary for achieving destructive interference
between the signal pathways in the upper and lower bus wave-
guides, thereby permitting transmission zeros to be realized in the
drop port response of the filter.

* An inverse Chebyshev filter is a filter with equi-ripples in the stopband and no


ripples in the passband. An elliptic or pseudo-elliptic filter is a filter with equi-
ripples in both the passband and the stopband. The transfer function of an ellip-
tic filter of order N has N poles and N transmission zeros, while a pseudo-elliptic
filter has only up to (N − 2) transmission zeros.
128 Optical Microring Resonators

Following the analysis of cascaded ADMRs in Section 3.5, we


define the transfer matrix of the first k stages of the microring
­ladder filter as

Tk = (Mk Pk )(Mk −1Pk −1 )(M2P2 )M1 , (3.100)

where Mk denotes the transfer matrix of the kth microring doublet


and Pk is the transfer matrix of the two connecting waveguides.
In terms of the round-trip delay variable, the transfer matrix Mk
of the symmetric microring doublet in stage k can be expressed in
the form

1  Fk ( z ) jK k z −1 
−1
Mk =  , (3.101)
Gk ( z −1 )  jK k z −1 Fk ( z −1 )

where z −1 = art e − jφrt and

K k = κ 12, k κ 2 , k , (3.102)

Fk ( z −1 ) = τ1, k − τ 2 , k (1 + τ12, k )z −1 + τ1, k z −2 , (3.103)

Gk ( z −1 ) = 1 − 2τ1, k τ 2 , k z −1 + τ12, k z −2 . (3.104)

In the above expressions, τ1, k = 1 − κ 12, k and τ 2 , k = 1 − κ 22 , k are the


transmission coefficients of the respective coupling junctions. The
transfer matrix Pk of the two bus waveguides of length L connect-
ing microring stages k and k + 1 with differential phase shift f­ actor
γk is given by

γ k 0
Pk = e − jβ b L  , (3.105)
0 1 

where βb is the propagation constant of the waveguides. Note that if


there is no differential phase shift (γk = 1), then Pk = e − jβ b L I, where I is
the identity matrix. Starting from T1 = M1 and using the recursive
relation Tk = (MkPk)Tk−1, we can show that the transfer matrix TN = T
of N cascaded stages can be expressed in the form

e − jβ b ( N −1)L  RN ( z −1 ) jσ N PN ( z −1 )z −1 
TN =  , (3.106)
QN ( z −1 ) −1 −1
 jPN ( z )z σ N RN ( z −1 ) 
Coupled Microring Optical Filters 129

where σN = γ2γ3…γN = ±1, and PN, RN, and QN are polynomials


­satisfying the recursive relations
Pk = γ k K k Rk −1 + Fk Pk −1 , (3.107)

Rk = γ k Fk Rk −1 − K k Pk −1 z −2 , (3.108)
k
Qk = Gk Qk −1 = ∏ Gn . (3.109)
n =1

The transfer functions at the drop port and through port of the micror-
ing ladder filter can be obtained from the transfer matrix T. Neglect­
ing the common phase factor e − jβ b ( N −1)L in Equation 3.106, we have
jPN ( z −1 )z −1
Hd ( z −1 ) = T21 = , (3.110)
QN ( z −1 )

RN ( z −1 )
H t ( z −1 ) = T11 = . (3.111)
QN ( z −1 )

Starting with the first microring stage with P1 = K1, R1 = F1, and
Q1 = G1, we can deduce from the recursive relations in Equations
3.107 through 3.109 that PN is a polynomial of degree 2(N–1), while RN
and QN are polynomials of degree 2N. Furthermore, since Kk and Fk
are even-degree and self-para-conjugate polynomials,* both PN and
RN are also of even-degree and self-para-conjugates. This implies
that the roots of PN and RN appear in both complex conjugate pairs

and para-conjugate pairs, for example, as {zk , 1/zk∗ } and {zk , 1/zk }.†
Note that a pair of conjugate roots located on the unit circle also sat-
isfies this property. Thus a microring ladder filter with N stages can
realize a drop port transfer function with 2N poles and up to 2(N–1)
transmission zeros that appear in complex and para-conjugate pairs.
Each microring doublet in the array is responsible for generating a
pair of complex conjugate poles in the transfer function.
Microring ladder filters can be synthesized using a procedure
based on the order reduction technique similar to the synthesis
* A polynomial P(z−1) of degree N is self-para-conjugate if P( z −1 ) = ± P ( z −1 ). The
polynomial is even self-para-conjugate if the coefficients of the kth and (N − k)th
power terms are equal; it is odd self-para-conjugate if these coefficients are equal
in magnitude but have opposite signs.
† In the s-domain, the transfer functions at the drop port and through port of the
microring ladder filter have transmission zeros that are quadrantally symmet-
ric, i.e., they appear as {± z, ±z*} (Liew and Van 2008).
130 Optical Microring Resonators

method developed for serially coupled microring filters in Section


3.4.4 (Liew and Van 2008). Suppose the target filter responses are
described by the transfer functions Hd and Ht of the form in Equations
(3.110) and (3.111), where PN is a polynomial of degree 2(N–1), and RN
and QN are polynomials of degree 2N. The roots of PN and RN appear
in conjugate pairs that either lie on the unit circle or are also paired
by their para-conjugates. Many common filters with symmetric spec-
tral responses, such as pseudo-elliptic and inverse Chebyshev filters,
satisfy the first property, that is, their transmission and reflection
responses have transmission nulls that are ­symmetrically located
about the center frequency. The second property (para-conjugate
zeros) is typically satisfied by filters with linear phase response.
The transfer functions Hd and Ht specified above can be real-
ized by an array of N microring doublets. The synthesis procedure
begins with the determination of the coupling parameters of the
last microring stage N, and proceeds backward until the first stage
is reached. Each microring doublet in stage k is completely charac-
terized by a pair of conjugate poles { pk , pk∗ } in the transfer functions.

Specifically, for each stage k, we select a pair of roots { pk , pk } from
the polynomial Qk to be the poles of the microring doublet. Since
these poles are the roots of the polynomial Gk in Equation 3.104,
they are given by

{ pk , pk∗ } = (τ 2 , k ± jκ 2 , k )/τ1, k . (3.112)

From the above expression, we can solve for the transmission coef-
ficients τ1,k and τ2,k of the microring doublet to get τ1,k = |pk|−2 and
τ2,k = τ1,kRe{pk}. Note that since the poles of a filter are typically
located in the right half of the (inverse) z-plane near the real axis, the
transmission coefficients τ1,k and τ2,k obtained from these formulas
are always positive. Knowledge of τ1,k and τ2,k allows us to construct
the transfer matrix Mk of stage k as given by Equation 3.101, which
is then de-embedded from the array. The transfer matrix Tk−1 of the
remaining k−1 stages is obtained by solving for the polynomials Pk−1
and Rk−1 from Equations 3.107 and 3.108. The results are

Fk Pk − K k Rk
Pk −1 = , (3.113)
Gk G k

Fk Rk + K k Pk z −2
Rk −1 = , (3.114)
γ k Gk G k
Coupled Microring Optical Filters 131

where G k = τ12, k − 2τ1, k τ 2 , k z −1 + z −2 is the Hermitian conjugate of Gk.


The differential phase factor γk of stage k is chosen to be either 1

or –1 such that Rk−1(pk−1)/Pk−1(pk−1) = j, where { pk −1 , pk −1 } is the pair of
poles selected for the next stage, k−1. The procedure is then repeated
with the parameter extraction of stage k−1 until the first stage is
reached.
Since the transfer matrices M k in Equation 3.101 are symmet-
ric with equal diagonal elements, they commute with each other,
MkMk−1 = Mk−1Mk. Thus, two microring doublets connected by
two parallel waveguides with no differential phase shift can be
interchanged without affecting the response of the filter. On the
other hand, if two microring doublets are connected by two
­waveguides with a differential π-phase shift, then it can be shown
that MkPM k−1 = P(M k−1PM k)P, where P = diag[−1, 1]. In this case the
order of the two stages can be exchanged if a π-phase shift is also
added before and after the two stages. These commutative proper-
ties of the matrices M k imply that the order in which the roots of
QN are used to determine the coupling parameters of the microring
stages does not affect the final filter design, except for a possible
permutation of the stage order and a different distribution of the
phase shift elements in the array. Moreover, it can also be shown
that by applying the above commutative properties, one can always
reduce the number of π-phase shifts in a given microring ladder
filter to at most one. A procedure for performing this phase shift
reduction is given in Prabhu et al. (2008b).
As an example, we consider the design of an eighth-order
pseudo-elliptic filter with a 100 GHz bandwidth, 0.5 dB in-band
ripple and −60 dB out-of-band rejection. Using a suitable filter
approximation technique (e.g., Martinez and Parks 1978), we obtain
a transfer ­function with eight poles and six zeros of the form given
by Equation 3.110 which satisfy these specifications. The roots of the
polynomials PN and QN are listed in Table 3.1. Using the Feldtkeller
relation in Equation 3.29, we also obtain the transfer function for the
reflection response of the form in Equation 3.111. The roots of the
polynomial RN are also given in Table 3.1. A plot of the pole-zero dia-
gram of the filter in the inverse z-plane is shown in Figure 3.19a and
the target spectral responses are shown in Figure 3.19b. We synthe-
size the transfer functions using a cascaded array of N = 4 microring
doublets. Assuming that the microring resonators have an FSR of
0.5 THz, we compute the transmission coefficients for each microring
132 Optical Microring Resonators

Table 3.1 Poles and Zeros of an Eighth-Order


Pseudo-Elliptic Filter
Transmission Zeros Reflection Zeros Poles
(Roots of PN) (Roots of RN) (Roots of QN)
0.3435 ± j0.9391 0.8129 ± j0.5825 0.7855 ± j0.5851
0.7238 ± j0.6901 0.8466 ± j0.5323 0.7691 ± j0.5161
0.6593 ± j0.7519 0.9159 ± j0.4013 0.7609 ± j0.3709
0.9879 ± j0.1554 0.7608 ± j0.1370

doublet and list their values in Table 3.2 along with the phase shift
factors. The factor γk = −1 in stage 3 indicates that there is a π-phase
shift between microring doublets 2 and 3. The spectral responses of
the microring ladder filter are shown in Figure 3.19b, which shows
good agreement between the target and synthesized responses.
Experimentally, a fourth-order microring ladder filter consist-
ing of two cascaded microring doublets has been demonstrated in
the SOI material (Masilamani and Van 2012). The filter was designed
to realize a fourth-order pseudo-elliptic transfer function with two
transmission zeros. Although the device exhibited a fourth-order
skirt roll-off, fabrication errors caused the zeros to be displaced
from their designed locations so that the measured filter response
did not show deep transmission nulls associated with the zeros.
Nevertheless, the microring ladder configuration holds promising
potential for realizing high-order optical filters with sharp band
transitions because they can be constructed and optimized stage
by stage, with each stage consisting of a simple microring doublet.
Another important advantage of the ladder configuration is that it
does not require negative coupling coefficients to realize transfer
functions with transmission zeros located on the unit circle (or on
the jω-axis in the s-domain), as in the example above. This is in

Table 3.2 Transmission Coefficients of the


Microring Ladder Filter with Four Stages
τ1,k τ2,k γk
Stage 1 0.7731 0.9842
Stage 2 0.9263 0.8304 1
Stage 3 0.8464 0.8989 −1
Stage 4 0.9795 0.8019 1
Coupled Microring Optical Filters 133
(a) 1 (b)
Poles 0
Transmission zeros
|Ht|2

Transmission (dB)
0.5 Reflection zeros
–20
Im{z}

0 –40
|Hd|2
–60
–0.5
–80
–1
–1 –0.5 0 0.5 1

50
00
50
00
0
0
50

0
0
0
0
–5

10
15
20
25
–2
–2
–1
–1
Re{z}
Frequency detune, ∆f (GHz)

Figure 3.19 (a) Pole-zero diagram of an eighth-order pseudo-elliptic filter


and (b) target filter responses (gray dashed lines) and spectral responses
of the synthesized microring ladder filter consisting of four cascaded
microring doublets (black solid lines).

contrast to filters based on 2D microring coupling topologies, which


are discussed in the next section.
An alternative filter structure that is closely related to the micror-
ing ladder filter is a Mach–Zehnder interferometer (MZI) whose
arms are loaded with APMRs (Madsen 1998). These microring-loaded
MZI structures (or RMZIs) are the photonic realizations of a general
class of digital filters known as sum–­difference all-pass filters, which
can realize Nth-order transfer functions containing up to N−2 trans-
mission zeros. In fact, it can be shown that an array of N cascaded
microring ­doublets is ­equivalent to an MZI with each arm loaded
with N APMRs (Van 2009). A significant difference between these
two structures is that the microring resonators in the RMZI filter gen-
erally have non-zero phase detunes, whereas they are synchronously
tuned in the microring ladder filter.

3.7 2D Networks of CMRs


CMR networks of 2D coupling topology are extensions of serially
coupled microring filters, which can be regarded as microring net-
works of 1D coupling topology. In Section 3.4, we showed that the
drop port transfer function of a serially coupled microring filter is
restricted to those with all poles and no transmission zeros. This
is due to the fact that near a resonant frequency, all signal path-
ways through the 1D microring array are nearly in phase, so it is
not possible to achieve destructive interference of the signals at the
134 Optical Microring Resonators

drop port to produce transmission nulls. In contrast, in a 2D CMR


network, light follows different pathways through the device which
can interfere destructively with each other at the output to produce
transmission nulls at certain frequencies. As a result, 2D micror-
ing coupling topologies allow more complex transfer functions to
be realized than possible with the 1D topology. Examples of filters
that can be synthesized with 2D CMR networks include inverse
Chebyshev and pseudo-elliptic filters with sharp band transitions,
and linear phase filters with constant group-delay response.
In this section we will first develop techniques for analyzing
and designing 2D CMR filters based on the energy coupling formal-
ism. This will be followed by an analysis of broadband or strongly
coupled microring networks in terms of power coupling. Important
differences in the spectral characteristics of the device obtained
from the two methods will also be highlighted.

3.7.1 Energy coupling analysis of 2D CMR networks


We consider a general network of N direct-coupled microring res-
onators arranged in a 2D square lattice as shown in Figure 3.20.
Other 2D arrangements of microrings are also possible but we will
focus on the 2D square lattice because it is the simplest coupling
topology to implement in practice. In general, there is no restriction
on the coupling ­topology except that it should not give rise to cou-
pling between counter-propagating modes in the microrings. This
precludes device configurations that contain coupling loops with an

si µi µ12 µ23

1 2 3 m
st
µ1, 2m

2m 2m–1 2m–2 m+1

N N–1 N–2
sd
µo

Figure 3.20 Schematic of a network of N microring resonators in a


2D square lattice coupling topology.
Coupled Microring Optical Filters 135

odd number of microrings, such as the triplet shown in Figure 3.21b.


In such a loop configuration, coupling between counter-propagat-
ing modes gives rise to a back-reflected wave at the input port of the
device, which is undesirable for most integrated optics applications.
We label the microrings in the network from 1 to N as shown
in Figure 3.20, although the order of numbering does not affect the
analysis that follows. For simplicity we assume that the microrings
have the same intrinsic decay rate, γ = αvg/2, due to propagation loss
α in the waveguides. However, we allow the microrings to have non-
identical resonant frequencies ωi to account for possible resonance
­mismatches among them. Coupling between two ­adjacent microrings
i and j is denoted by the energy coupling coefficient μi,j. Microrings 1
and N are also coupled to the input and output waveguides via input
and output coupling coefficients μi and μo, respectively.
We denote the wave amplitude in each microring i as ai, which
is normalized so that |ai|2 gives the energy stored in the resonator.
The signals si, st, and sd in the input and output waveguides denote
the input, through, and drop signals, respectively. These signals
represent the rates of energy supplied to or drawn from the net-
work and are thus normalized with respect to power. By examining
the coupled mode equations in Equation 3.38 for a serially coupled
microring filter, we can generalize them for a 2D CMR n ­ etwork as
follows (Van 2007):

 a1  ( jω1 − γ 1 ) − jµ1, 2 − jµ1, 3  − jµ1, N   a1  − jµ i si 


 a   − jµ ( jω 2 − γ ) − jµ 2 , 3  − jµ 2 , N   a2   0 
2 1, 2
d       
 a3  =  − jµ1, 3 − jµ 2 , 3 ( jω 3 − γ )  − jµ 3 , N   a3  +  0  .
dt        
           
aN   − jµ1, N − jµ 2 , N − jµ 3 , N  ( jω N − γ N ) aN   0 
(3.115)

(a) (b)

1 2
3

4 3 1 2

Figure 3.21 Examples of unrealizable or undesirable microring coupling


topologies: (a) a quadruplet with cross couplings and (b) a triplet that
causes coupling between counter-propagating modes in the microrings.
136 Optical Microring Resonators

The above equation assumes the most general coupling t­opology


where each microring i is coupled to every other microring j in the
network via coupling coefficient μi,j. If there is no coupling between
microrings i and j, then μi,j = 0. The amplitude decay rates γ1 and
γN represent the total decay rates in resonators 1 and N, respec-
tively, due to both intrinsic loss and coupling to the input or output
waveguide,

γ (1,N ) = γ + γ (i,o) , (3.116)


2
where γ (i,o) = µ(i,o) /2.
To obtain the frequency response of the device, we assume a
harmonic input excitation of the form si ~ ejωt. The signals in the
­microrings will also have the same time dependence, ai ~ ejωt. The
input frequency ω is detuned from the resonant frequency of
microring i by an amount given by

ω − ω i = (ω − ω 0 ) − (ω i − ω 0 ) = ∆ω − δω i , (3.117)

where δωi is the detune of the resonant frequency of microring i


from a center frequency ω0 (typically chosen as the center frequency
of the filter passband). Substituting the above expressions into
Equation 3.115, we can write the resulting equation in matrix form as
(sI + L + jM)a = s, (3.118)

where s = jΔω + γ, a = [a1, a2…aN−1, aN]T, s = [−jμisi, 0, …0, 0]T, and I is


the N × N identity matrix. The matrix L is a diagonal matrix that
accounts for couplings to the input and output waveguides,

L = diag[γ i , 0, … 0, γ o ], (3.119)

and M is an N × N symmetric coupling matrix,

 δω1 µ 1, 2 µ 1, 3  µ 1, N 
µ δω 2 µ 2,3  µ 2, N 
 1, 2 
M =  µ 1, 3 µ 2,3 δω 3  µ3,N  . (3.120)
 
      
µ1, N µ 2, N µ3,N  δω N 

For 1D CMR (or CROW) filters, the coupling matrix M has a simple
tridiagonal form.
Coupled Microring Optical Filters 137

The solution of the matrix equation (3.118) can be expressed


in closed form. We diagonalize the matrix −(L + jM) in the form
−(L + jM) = Q ⋅ D ⋅ Q−1, where D is a diagonal matrix containing the
eigenvalues of −(L + jM) and Q is a matrix containing the corre-
sponding eigenvectors. Making this substitution in Equation 3.118
and solving for a, we get

a = Q(sI − D)−1 Q −1s. (3.121)

The wave amplitude ai can be explicitly expressed as


N
Qi , k Qk−,11
ai = − jµ i si ∑k =1
s − pk
, i = 1 to N , (3.122)

where pk is the kth diagonal element of D, and Qi,k and Qk−,11 are the
matrix elements of Q and Q−1, respectively. Using the relations
st = si − jμia1 and sd = −jμo aN, we obtain the following expressions for
the transfer functions at the through port and drop port of the CMR
network:
N
Q1, k Qk−,11 R(s)
H t (s) = 1 − µ i2 ∑
k =1
s − pk

Q(s)
, (3.123)

N
QN , k Qk−,11 P(s)
Hd (s) = −µ i µ o ∑
k =1
s − pk

Q(s)
. (3.124)

The above equations show that the 2D CMR filter has N poles
which, in the absence of loss (γ = 0), are given by the eigenvalues pk of
the matrix −(L + jM). The numerator polynomial R(s) of the through
port transfer function has degree N, while the numerator polyno-
mial P(s) of the drop port transfer function has a maximum degree of
N − 2. The latter result follows from the fact that the coefficient of the
highest power term of P(s)—that is, the (N−1)th power term—is zero,
N

∑Q
k =1
N ,k Qk−,11 = 0,

since it is the product of row N of Q and column 1 of Q−1. Thus a


2D network of N microrings can realize an Nth-order filter transfer
function containing up to N − 2 transmission zeros.
138 Optical Microring Resonators

If we take the absolute square of each term in the summation in


Equation 3.124, the result has the form of a Lorentzian resonance:
2
QN , k Qk−,11 1
µiµo ∝ , (3.125)
j(ω − ω 0 ) + γ − pk (ω − Ω k )2 + Γ 2k

where Ωk = ω0 − Im{pk} and Γk = γ + Re{pk}. Thus the spectral response


of a 2D CMR network can be regarded as consisting of a sum of
Lorentzian resonances centered at frequencies Ωk with linewidths
Γk.* The Lorentzian response is characteristic of the weak coupling
approximation assumed in the energy coupling formalism.

3.7.2 Energy coupling synthesis of 2D CMR networks


We consider next the problem of synthesizing a 2D CMR network
to achieve a target filter response. The synthesis problem was first
considered in Van (2007) in which the CMR network was modeled
by an equivalent electrical circuit. By applying a suitable electri-
cal filter synthesis method (such as in Atia et al. 1974), the micror-
ing coupling parameters could be determined for a given transfer
function. In this section, we will develop a simpler and more direct
method for synthesizing 2D CMR networks without making use of
an equivalent electrical circuit.
We consider a lossless† 2D CMR filter consisting of N micror-
ing resonators with identical resonant frequency ω0. Setting γ = 0
in Equation 3.115, we rewrite the coupled mode equations for the
CMR in the form

 a1   jω 0 − jµ1, 2 − jµ1, 3  − jµ1, N   a1   u1 


 a   − jµ jω 0 − jµ 2 , 3  − jµ 2 , N   a2   0 
d       
2 1, 2

 a3  =  − jµ1, 3 − jµ 2 , 3 jω 0  − jµ 3 , N   a3  +  0  ,
dt       
              
 aN   − jµ1, N − jµ 2 , N − jµ 3 ,N  jω 0   aN  uN 
(3.126)
* This statement is accurate if the resonances are far apart. When the resonances
are close together or overlap, coupling between them gives rise to a spectral
response that drastically deviates from the Lorentzian shape.
† For CMR filters with uniform microring loss, the transfer functions can be
predistorted to compensate for the effect of loss before applying the synthesis
­procedure. The predistortion technique for synthesizing microring filters with
loss can be found in Prabhu and Van (2008).
Coupled Microring Optical Filters 139

where

 jµ i 
u1 = − γ i a1 − jµ i si = − jµ i  si − a1 , (3.127)
 2 

 jµ o 
uN = − γ o aN − jµ o sa = − jµ o  sa − aN  . (3.128)
 2 

In Equation 3.128 sa is the signal applied to the add port of the fil-
ter. The quantities |u1|2 and |uN|2 represent the net powers sup-
plied to microrings 1 and N, respectively, from the input and
output ­waveguides. Assuming signals with harmonic time depen-
dence ejωt, we can express Equation 3.126 as

(sI + jM)a = u, (3.129)

where s = j(ω − ω0), u = [u1, 0, …0, uN]T, and M is the coupling


matrix in Equation 3.120 with zero frequency detunes,

 0 µ 1, 2 µ 1, 3  µ 1, N 
µ 0 µ 2,3  µ 2, N 
 1, 2 
M =  µ 1, 3 µ 2,3 0  µ3,N  . (3.130)
 
      
µ1, N µ 2, N µ3,N  0 

Since M is real and symmetric, we can diagonalize it in the form


M = W ⋅ Λ ⋅ WT, where Λ is a diagonal matrix containing the eigen-
values and W is a real unitary matrix containing the eigenvectors.
Making this substitution in Equation 3.129 and solving for a, we get

Λ )−1 W T u.
a = W(sI + jΛ (3.131)

Specifically, a1 and aN are given by


N N
W12, k W1, kWN , k
a1 = u1 ∑
k =1
s + jλ k
+ uN ∑
k =1
s + jλ k
≡ A11u1 + A1N uN , (3.132)

N N
WN , kW1, k WN2 , k
aN = u1 ∑
k =1
s + jλ k
+ uN ∑k =1
s + jλ k
≡ AN 1u1 + ANN uN , (3.133)
140 Optical Microring Resonators

where λk are the eigenvalues of M and Wi,j are the elements of


the matrix W. If we remove the output waveguide so that the CMR
becomes an all-pass network, then μo = 0, which also gives uN = 0.
Under this condition, we obtain from Equations 3.132 and 3.133 the
following expressions for the network parameters A11 and AN1:
N
a1 W12, k
A11 =
u1 µo = 0
= ∑
k =1
s + jλ k
, (3.134)

N
a WN , kW1, k
AN 1 = N
u1 µo = 0
= ∑
k =1
s + jλ k
. (3.135)

Next we establish the relationships between the parameters


A11 and AN1 and the specified drop port and through port transfer
­functions of the filter. The input and output signals of the CMR
­filter are related by a scattering matrix S defined by

 st  S11 (s) S12 (s)  si 


 s  = S (s) S22 (s)  sa 
. (3.136)
 d   21
From Equations 3.127 and 3.128, we can express si and sa in terms of
u1 and uN as

1  µ i2 
si = − u + a1 , (3.137)
jµ i  2 
1

1  µ o2 
sa = − u + aN  . (3.138)
jµ o 
N
2 
At the input and output bus coupling junctions we also have
st = si − jμia1 and sd = sa − jμo aN. Substituting the expressions for si
and sa in Equations 3.137 and 3.138 into these relations, we get

1  µ i2 
st = −  u − a1 , (3.139)
2 
1
jµ i

1  µ o2 
sd = − u − aN  . (3.140)
jµ o 
N
2 

Equations 3.137 through 3.140 link the input and output signals
si, sa, st, and sd of the CMR network to the newly defined signals
Coupled Microring Optical Filters 141

u1 and uN. Substituting these expressions into Equation 3.136 and


­solving for a1 and aN, we obtain

2  1 − S11 + S22 − ∆ S  4  S12 


a1 = 2   u1 − uN , (3.141)

µ i 1 + S11 + S22 + ∆ S  µ i µ o 1 + S11 + S22 + ∆ S 

4  S21  2  1 + S11 − S22 − ∆ S 


aN = −   u1 + 2  uN ,
µ i µ o  1 + S11 + S22 + ∆ S  µ o  1 + S11 + S22 + ∆ S 
(3.142)
where ΔS = S11S22 − S12S21 is the determinant of the scattering matrix.
Comparing the above equations to Equations 3.132 and 3.133, we get

a1 2  1 − S11 + S22 − ∆ S 
A11 = = , (3.143)
u1 uN = 0 µ i2  1 + S11 + S22 + ∆ S 

aN 4  S21 
AN 1 = =− . (3.144)
u1 uN = 0 µ i µ o  1 + S11 + S22 + ∆ S 

The above equations allow us to compute network parameters A11


and AN1 from the scattering parameters of the CMR filter.
Since an Nth-order CMR filter is a linear passive network,
its scattering matrix has the general form

1  R(s) σjP(− s)


S= , (3.145)
Q(s)  jP(s) σR(− s) 

where σ = +1/−1 for N even/odd. The polynomials R and Q have


degree N, while P has maximum degree N−2. The determinant of
S is given by

σ [ R(s)R(− s) + P(s)P(− s)] Q(− s)


∆S = 2
=σ , (3.146)
Q (s) Q(s)

where we have made use of the Feldtkeller relation in Equation 3.26.


By writing the elements of the scattering matrix in Equations 3.143
and 3.144 in terms of the polynomials P, R, and Q, we obtain the
­following expressions for A11 and AN1:

2  Q(s) − R(s) + σR(− s) − σQ(− s)  M(s)


A11 =  Q(s) + R(s) + σR(− s) + σQ(− s)  ≡ D(s) , (3.147)
µ i2  
142 Optical Microring Resonators

4  jP(s)  4 jP(s)
AN1 = −   =− .
µ i µ o  Q(s) + R(s) + σR(− s) + σQ(− s)  µ i µ o D(ss)
(3.148)

In the above expressions, the polynomials M(s) and D(s) have maxi-
mum degree N with the leading coefficient of D(s) assumed to be 1.
Equations 3.147 and 3.148 allow us to determine the parameters A11
and AN1 of the CMR network from the specified transfer functions
of the filter.
In the synthesis procedure of a 2D CMR filter, given the
drop port and through port transfer functions of the filter of
the form

jP(s) j( pN − 2 s N − 2 + pN − 3 s N − 3 +  + p1s + p0 )
Hd (s) = = , (3.149)
Q(s) s N + qN −1s N −1 +  + q1s + q0

R(s) rN s N + rN −1s N −1 +  + r1s + r0


H t (s) = = , (3.150)
Q(s) s N + qN −1s N −1 +  + q1s + q0

we can determine the coupling coefficients of the CMR network as


follows. Since the eigenvector matrix W is a unitary matrix, its rows
have unity magnitude. Specifically, for row 1 we have

∑W
k =1
2
1, k = 1. (3.151)

From Equation 3.134 we recognize that the above expression is


also the coefficient of the (N–1)th power term of the numerator
polynomial M(s) of A11. Thus, by setting this coefficient to 1 in
the numerator polynomial of Equation 3.147 and solving for μi,
we obtain

2(qN −1 − rN −1 )
µ i2 = . (3.152)
1 + rN

For filter designs based on symmetric CMR networks (i.e., S11 = S22),
we can set μo = μi.
Knowledge of μi allows us to determine the polynomials M(s)
and D(s) of A11 as defined in Equation 3.147. Next, performing
Coupled Microring Optical Filters 143

partial fraction expansions of the rational functions M/D and P/D,


we can express the results as
N
M(s) ξ(k11)
D(s)
= ∑
k =1
s − ρk
, (3.153)

N
P(s) ξ(k21)
D(s)
= ∑k =1
s − ρk
, (3.154)

where ρk are the poles and ξ(k11) and ξ(k21) are the residues of the
respective rational function. Comparing the above expressions to
Equations 3.134 and 3.135 shows that the eigenvalues of the matrix
M are given by λk = jρk. The elements W1,k and WN,k of the first and
last rows of the matrix W are computed from the residues ξ(k11) and
ξ(k21) as follows (Cameron 1999):

1/2
W1, k = ξ(k11) , (3.155)

{ }
WN , k = sgn Im ξ(k21)  W1, k , (3.156)

where the sgn function returns the sign of its argument. The remain-
ing rows of the matrix W can be obtained by Gram–Schmidt ortho-
normalization. Finally the energy coupling matrix M is obtained
from M = WΛWT, where Λ is the diagonal matrix containing the
eigenvalues λk.
The above synthesis procedure typically yields a full coupling
matrix M which corresponds to a coupling topology that may
not be realizable due to physical layout constraints. For example,
the coupling topology may require a microring to be coupled
to too many other microrings so that adjacent resonators would
touch or overlap each other. It may also contain quadruplets with
cross-couplings as shown in Figure 3.21a, or coupling loops with
an odd number of microrings as shown in Figure 3.21b. The lat-
ter ­structures are undesirable since they lead to coupling between
counter-propagating modes in the microrings and result in a
reflected wave at the input port. In general, it is possible to con-
vert an unrealizable—or undesirable—coupling configuration to
a simpler and realizable one by applying similarity transforma-
tions such as Jacobi rotations to the coupling matrix M without
144 Optical Microring Resonators

disturbing its eigenvalues. Each Jacobi rotation yields a new cou-


pling matrix M′ according to

M′ = R(θr ) ⋅ M ⋅ R T (θr ), (3.157)

where R(θr) is an N × N rotation matrix and θr is the rotation angle


chosen to annihilate an undesirable coupling element in the origi-
nal coupling matrix M (Cameron 1999, 2003). We also require that
the first and last rows of R to have 1s on the diagonal and zeros
as the off-diagonal elements so as not to disturb the first and last
microrings, since these are fixed by the input and output port loca-
tions. In general, we can always reduce a given coupling matrix to
the so-called 2 × m canonical form, which has the fewest number of
coupling elements. The canonical coupling topology has the form of
a folded microring array, such as the one shown in Figure 3.22a for a
six-ring network. A procedure for reducing a given coupling matrix
M to the canonical coupling form can be found in Cameron (1999).
In practical microring filter design, we typically perform the
CMR synthesis procedure for a prototype filter with normalized
center frequency ω0 = 1 rad/s and cutoff frequency ωc = 1 rad/s
(bandwidth of 2 rad/s). The energy coupling coefficients are then
scaled to obtain a filter with a specified bandwidth B (rad/s) accord-
ing to the relations µ ( i ,o ) = µ( i ,o ) B/2 and µ i , j = µ i , j (B/2), where
the tilde sign denotes parameters of the new bandpass filter with
bandwidth B. In the physical implementation of the microring filter,
the energy coupling coefficients are converted to the field coupling

(a) (b) 0

–10
|Ht|2
Transmission (dB)

si µi µ12 µ23
–20
1 2 3 –30
st |Hd|2
sd µ16 µ25 µ34 –40

6 5 4 –50

–60
µi µ56 µ45
–70
–4 –3 –2 –1 0 1 2 3 4
Normalized frequency detune, ∆ω/(B/2)

Figure 3.22 (a) Schematic of a 6th-order canonical CMR filter and


(b) ­target filter responses (gray dashed lines) and spectral responses of
the synthesized canonical CMR filter (black solid lines).
Coupled Microring Optical Filters 145

coefficients for microring resonators with a specified FSR (in Hertz)


using the relations κ ( i ,o ) = µ ( i ,o ) / FSR and κ i , j = µ i , j /FSR.
We illustrate the above synthesis procedure with the design of
a sixth-order pseudo-elliptic optical filter with 0.05 dB ripple in the
passband and −40 dB stopband rejection. Choosing a design with
four transmission zeros, we employ a suitable filter approximation
method (e.g., Ellis 1994) to obtain a sixth-order transfer function that
meets the above specifications. The polynomials of the drop port
and through port transfer functions of the prototype filter (with a
bandwidth of 2 rad/s) are given by
P(s) = 0.1128(s 4 + 4.5861s2 + 4.9410),
R(s) = s6 + 1.6716s 4 + 0.7452s2 + 0.0589,
Q(s) = s6 + 1.9528s5 + 3.5783s 4 + 3.8036s3 + 3.1742s2
+ 1.7013s + 0.5606. (3.158)
The filter has six poles located at pk = {−0.0744 ± j1.0609, −0.2961 ±
j0.9154, −0.6059 ± j0.4103}. The drop port transfer function has four
transmission zeros located on the imaginary axis at zk = {± j1.6900,
±j1.3153}, while the through port transfer function has six zeros on
the imaginary axis at zk = {± j0.3166, ± j0.7834, ± j0.9786}. The target
spectral responses of the filter are shown by the gray dashed lines
in Figure 3.22b.
In the microring filter design, we first use Equation 3.152 to
compute the input and output coupling coefficients of the CMR fil-
ter to get μi = μo = 1.3974. Knowledge of μi and μo allows us to deter-
mine the polynomials M(s) and D(s) from Equations 3.147 and 3.148.
Performing partial fraction expansions of the rational functions
M/D and P/D, we obtain the poles and residues that are listed in
Table 3.3. The first and last rows of the orthonormal matrix W are

Table 3.3 Poles and Residues of the


Network Parameters A11 and AN1
Poles ρk ξ(k11) ξ(k21)
j1.1429 0.0752 −j0.1468
−j1.0507 0.0752 j0.1468
j1.0507 0.1677 j0.3275
−j1.0507 0.1677 −j0.3275
j0.4635 0.2571 −j0.5020
−j0.4635 0.2571 j0.5020
146 Optical Microring Resonators

next calculated using Equations 3.155 and 3.156 and the remaining
rows are obtained by Gram–Schmidt orthonormalization. From W
and the eigenvalue matrix Λ, we calculate the coupling matrix M
to be (only the upper half is shown since the matrix is symmetric)

0 −0.2175 0.5516 −0.5329 −0.1957 −0.0578 


 −0.8264 0.0740 0.1033 −0.5557 −0.2897 
 
 0.5528 −0.0439 −0.1381 0.4749 
M= .
 −0.4852 −0.1671 0.4712 
 0.7589 0.3774 
 
 0 
(3.159)

The above matrix corresponds to a coupling topology that is not


physically realizable since it requires every microring in the net-
work to be coupled to all the other microrings. By applying a series
of matrix rotations as described in Cameron (1999), we can reduce
the above coupling matrix to the canonical form

0 0.8209 0 0 0 0.0578 
 0 0.5393 0 0.2840 0
 
 0 0.7820 0 0
MC =  .
 0 −0.5393 0
 0 −0.8209 
 
 0 
(3.160)

It is evident that the canonical coupling matrix has much fewer


coupling elements than the original matrix in Equation 3.159.
A ­schematic of the canonical CMR filter is shown in Figure 3.22a. Its
spectral responses are shown by the black solid lines in Figure 3.22b,
which are in good agreement with the ­target filter responses.
We observe in Equation 3.160 that all the diagonal elements of the
canonical coupling matrix MC are zero, implying that the microring
resonators are synchronously tuned to the center frequency ω0 of
the filter passband. We also note that the coupling elements μ45 and
μ56 are negative. In general, 2D CMR networks require negative cou-
pling elements to realize transfer functions with transmission zeros
on the jω-axis. One way to realize a negative coupling element is
Coupled Microring Optical Filters 147
(a) (b) (c)
π

3π/2 < θc < 2π λ/8


(–)

Figure 3.23 Schemes for realizing negative couplings in a CMR network:


(a) micro-racetracks with coupling phase θc = kcL such that 3π/2 < θc < 2π,
(b) quadrupole coupling loop sheared by λ/8, and (c) cascaded microring
doublets with inter-stage π-phase shift.

by evanescent coupling between two micro-racetracks with a long


coupling length L such that 3π/2 < kcL < 2π, where kc is the coupling
strength, as shown in Figure 3.23a (Van 2007). Alternatively, nega-
tive coupling can be simulated in a loop-coupled microring qua-
druplet by shearing the loop by (2n + 1)λ/8, where n is an integer, as
shown in Figure 3.23b (Popovic 2007, Tsay and Van 2012b). Finally,
we note that the requirement for negative coupling elements is
­eliminated by using the microring ladder filter configuration in
Section 3.6. Indeed, it can be shown that any CMR network with
negative coupling coefficients can be decomposed into an array of
cascaded microring doublets with all-positive ­coupling coefficients
and differential π-phase shifts between the stages (Prabhu et al.
2008b). This is illustrated in Figure 3.22c for a quadrupole ­microring
filter with a negative coupling element.
Although 2D CMR filters are generally more difficult to ­realize
than serially coupled microring filters, attempts at demonstrat-
ing quadrupole 2D CMR filters with negative coupling using the
sheared coupling loop configuration have been reported in SOI
and SiN material systems (Popovic et al. 2008, Bachman et al. 2015).
In particular, the measured transmission response of the silicon
quadrupole pseudo-elliptic filter in Bachman et al. (2015) showed
steep roll-offs and deep extinctions due to transmission nulls, thus
providing clear evidence of negative coupling in the structure.

3.7.3 Power coupling analysis of 2D CMR networks


In the energy coupling analysis of 2D CMR networks, we implic-
itly implicitlyassume that the field amplitude in each microring
resonator is uniformly distributed around the microring. This
148 Optical Microring Resonators

assumption, which holds for weak coupling, allows us to simplify


the analysis by considering only the energy stored in each resona-
tor (given by |ai(t)|2). Since the energy coupling formalism does not
take into account the phase or amplitude variation of the signal in
the microring waveguide, the response of the CMR network does
not depend on the exact l­ocations of the coupling junctions on the
microrings.
For strongly coupled microring resonators, the field amplitudes
before and after a coupling junction can differ significantly, and
a more accurate analysis based on the coupling of power in space
must be used to describe the device response. Such an analysis will
also take into account the exact locations of the coupling junctions
on the microrings. In this section, we develop a general method
for analyzing 2D CMR networks using the power coupling formal-
ism. We will show that any 2D microring coupling topology can be
described in terms of a direct coupling matrix plus a commutator
matrix responsible for nonadjacent resonator couplings. The latter
term, which is neglected in the energy coupling approximation, can
give rise to prominent resonance features that are not observed in
weakly coupled CMR networks (Tsay and Van 2011b).
We consider again a 2D CMR network consisting of N micror-
ings arranged in a square lattice as shown in Figure 3.24a.
To ­simplify the analysis, we assume that the microrings are identi-
cal with the same radius R. Coupling between two adjacent micror-
ings i and j is denoted by the field coupling coefficient κi,j. The input
and output coupling coefficients are denoted by κi and κo, respec-
tively. The ­signals at the input, through, drop, and add ports are
labeled si, st, sd, and sa, respectively, which are normalized with
respect to power. In each microring i, we label the amplitude of the
circulating wave after each coupling junction as ai, bi, ci, di, in the
order of the direction of wave propagation, as shown in Figure 3.24a.
These waves are also normalized with respect to power. The label-
ing order, a → b → c → d, is also followed for signal paths crossing
into ­adjacent microrings. For example, in Figure 3.24a, the field
a1 in ring 1 is c­ oupled to field b2 in ring 2 via coupling coefficient
κ12, which is then coupled to field c2m−1 in ring 2m−1 via coupling
­coefficient κ2,2m−1, and so on. This order of field labeling yields
­coupling ­matrices with simple structures in the analysis below.
To facilitate the analysis of the CMR network, we transform the
structure into an equivalent array of coupled straight waveguides
(Tsay and Van 2011b). This is achieved by “cutting” each microring
Coupled Microring Optical Filters 149
(a)
si st
Cut point
κi κ12 κ23
a1 a2 a3 am
d1 d2
1 b 2 3 m
1 b2
c1 c2

κ1, 2m κ2, 2m–1

c2m–1
2m 2m–1 m+1
2m–2
a2m a2m–1

κN–1, N
cN

N bN N–1 N–2
dN
aN aN–1 aN–2
κo

sa sd

(b)
si st
κi z–1/4
a1' 1
a1 κ12 b1 c1 d1
a2' 2
a2 b2 κ23
a3' 3
a3
κ1, 2m κ2, 2m–1
2m–2
a2m–2
2m–1
a2m–1
2m
a2m

aN–2
N–2
aN–1
κN–2, N–1
N–1
aN κN–1, N
'
aN
κo N
sa sd
[M1] [M2] [M3] [M4]
[L]

Figure 3.24 (a) Schematic of a 2D CMR network consisting of N micror-


ing resonators and (b) equivalent unfolded coupled waveguide array of
the CMR structure in (a).

at the point just before the field ai is defined and unfolding it into
a straight waveguide while keeping track of the coupling junc-
tions between adjacent microrings. In this manner, the CMR struc-
ture in Figure 3.24a is transformed into an equivalent “unfolded”
­coupled-waveguides configuration shown in Figure 3.24b, where a
connection between two waveguides denotes a coupling junction.
150 Optical Microring Resonators

The unfolded structure resembles an array of coupled FP waveguide


cavities, except that the facets have periodic instead of reflective
boundary conditions so that the waves travel only in the forward
direction.
The transfer functions of the CMR network can now be obtained
by analyzing the unfolded coupled waveguide array using the trans-
fer matrix method. We can view the waveguide array as consisting
of four cascaded sections, each representing a phase delay of a quar-
ter ring, e − jφrt /4 = z −1/4, where ϕrt is the microring round-trip phase.
We denote the fields in each section by the arrays a, b, c, and d, where
T
a =  a1 , a2 ,  aN −1 , aN  , (3.161)

with similar expressions for b, c, and d. The sections are connected


to each other by a set of coupling junctions that can be described by
the four transfer matrices M1, M 2, M3, and M4. These N × N matrices
are symmetric with the property that if there is coupling between
waveguides i and j at junction k (k = 1–4) with coupling coefficient
κi,j, then
M k (i , i ) = M k ( j , j ) = τ i , j ,
(3.162)
Mk (i , j) = Mk ( j , i) = − jκ i , j ,

where τ i2, j + κ i2, j = 1. If waveguide i is uncoupled at junction k,


then M k(i,i) = 1. Specifically, with the field labeling as shown in
Figure 3.24a, the matrix M1 has the block-diagonal form

K12 
 K 34 
 
M1 =   , (3.163)
 
 Ki, j 
 

where Ki,j is the 2 × 2 coupling matrix associated with the coupling


junction between microrings i and j,

 τ i , j − jκ i , j 
Ki, j =  . (3.164)
 − jκ i , j τ i , j 

The other coupling matrices M k can also be transformed to the above


block-diagonal form through some suitable permutations. Using
Coupled Microring Optical Filters 151

these transfer matrices, we can relate the field arrays through the
expressions b = z−1/4M1a, c = z−1/4M 2 b, d = z−1/4M3 c, and a′ = z−1/4M4 d,
where a ′ is the field array defined just before the input and output
coupling junctions, as indicated in Figure 3.24b. These relations can
be combined to give

a′ = z −1M 4 M3 M2 M1a ≡ z −1Ma, (3.165)

where M represents the ring-to-ring coupling matrix of the CMR


network. At the left boundary of the coupled waveguide array, the
field arrays a and a′ are also related by

a = La′ + s, (3.166)

where s = [−jκisi, 0, …0, − jκosa]T is the input field array and L is a


diagonal matrix representing the bus-to-ring couplings,

L = diag  τ i , 1,  1, τ o  , (3.167)

2
with τ( i,o ) = 1 − κ ( i,o ) . Upon substituting Equation 3.165 into
Equation 3.166, we obtain

(I − z −1LM)a = s, (3.168)

where I is the N × N identity matrix. The above equation shows that


the spectral responses of the field amplitudes in a CMR network are
characterized by a field coupling matrix M, which is determined by
the coupling topology, and a ring-to-bus coupling matrix L represent-
ing extrinsic losses due to coupling to the bus waveguides. Internal
loss in the microrings can be accounted for by letting z −1 = art e − jφrt,
where art is the round-trip amplitude attenuation. We also note that
Equation 3.168 has a somewhat similar form to Equation 3.118 obtained
using the energy coupling formalism. Indeed, it can be shown that
Equation 3.118 can be derived from the more general Equation 3.168
under the weak coupling condition (Tsay and Van 2011b).
The field coupling equation in Equation 3.168 can be solved by
diagonalizing the matrix product LM in the form LM = QDQ−1,
where D is a diagonal matrix containing the eigenvalues of LM and
Q is the corresponding eigenvector matrix. Making this substitu-
tion in Equation 3.168 and solving for a, we get

a = Q(I − z −1D)−1 Q −1s. (3.169)


152 Optical Microring Resonators

Assuming that only the input port is excited (si ≠ 0, sa = 0), we


obtain from Equation 3.169 the field amplitude ai in microring i as
follows:
N
Qi , k Qk−,11
ai = − jκ i si ∑
k =1
1 − z −1
λ k
, (3.170)

where λk are the diagonal entries of D, and Qi,k and Qk−,11 are elements
of the matrices Q and Q−1, respectively.
The transfer functions at the through port and drop port of the
CMR network can be obtained by relating the output signals st and
sd to the fields a1 and aN in microrings 1 and N. At the coupling junc-
tion between the input waveguide and microring 1, we have the
relations

st = τ i si − jκ i a1′ , (3.171)

a1 = − jκ i si + τ i a1′ . (3.172)

Eliminating a1′ from the above equations gives

1
st = (si − jκ i a1 ). (3.173)
τi

Using Equation 3.170 for a1, we obtain the through port transfer
function of the CMR filter as follows:

st 1  Q1, k Qk−,11 
N
H t ( z −1 ) = =  1 − κ i2
si τ i  k =1

1 − z −1λ k 
. (3.174)

To determine the drop port response, we set sa = 0 so that at the


output coupling junction of microring N, we have sd = − jκ o aN′ and
aN = τ o aN′ . These two relations are combined to give sd = −jκo aN/τo.
Using Equation 3.170 for aN, we obtain, for the drop port transfer
function of the CMR filter,
N
s κκ QN , k Qk−1,1
−1
Hd ( z ) = d = − i o
si τo ∑k =1
1 − z −1λ k
. (3.175)

Equations 3.174 and 3.175 show that, in the absence of reso-


nator loss, the poles of the CMR filter are given by the inverse
Coupled Microring Optical Filters 153

of the eigenvalues (1/λk) of the matrix product LM. We also note


that the zeroth-order term of the numerator polynomial of Hd is
zero,

κ iκ o N

τo
∑ QN,kQk−,11 = 0, (3.176)
k =1

since the summation is the product of row N of Q and column 1


of Q−1. Thus we can write the drop port transfer function in the
form

z −1P( z −1 )
Hd ( z −1 ) = , (3.177)
Q( z −1 )

where P is a polynomial of maximum degree N−2. Equation 3.177


indicates that excluding the trivial zero at z−1 = 0, the drop port trans-
fer function of the CMR network has a maximum of N−2 transmis-
sion zeros. We recall that this was also shown to be the case for the
device transfer function in the s-domain (Equation 3.124).
If we take the absolute square of each term in the summation of
Equation 3.175, we obtain an expression of the form
2
QN,k Qk−,11 1
∝ , (3.178)
1 − λ k e − jφrt 1 + Fk sin 2 (φ)

where φ = (ϕrt − θk)/2, θk is the angle of λk, and Fk = 4|λk|/(1 − |λk|2).


The above expression has the form of an Airy function. Thus the
response of a CMR network can be regarded as consisting of a sum
of N Airy resonances.* In contrast, energy coupling analysis shows
that the response of a weakly coupled CMR network consists of a
sum of N Lorentzian resonances.
We also saw in Section 3.7.1 that in the energy coupling for-
malism, the coupling matrix M of a 2D CMR network specifies
the direct couplings between adjacent microring resonators, that
is, the element M(i, j) specifies the energy coupling coefficient μi,j
between neighbor microrings i and j. In contrast, the structure of
the complex coupling matrix M in the power coupling formalism
is more complicated. To deduce the physical meaning of the field
coupling matrix M, we first look at the individual coupling matrices

* See also Footnote on page 138.


154 Optical Microring Resonators

Mk of the coupled waveguide array. Each of these matrices, through


some suitable permutation, can be expressed in the block-diagonal
form of Equation 3.163, in which the block matrices are the junc-
tion coupling matrices Ki,j. Since Ki,j are circular matrices,* the cou-
pling matrices M k are also circular, which means that they can be
expressed in the form

Mk = Wk Λ k WkT = e jΨk , (3.179)

where Λk is a diagonal matrix containing the eigenvalues of M k and


Wk is a real matrix containing the corresponding eigenvectors. The
eigenvalues of Mk are given by
jθi , j
Λ k (i , i) = Λ*k ( j , j) = τ i , j − jκ i , j = e , (3.180)

where θi,j = −tan−1(κi,j/τi,j) is defined as the coupling angle between


waveguides i and j. If waveguide i of section k is not coupled to any
other waveguide, then Λk(i,i) = 1. The matrix Ψk in Equation 3.179 is
real and symmetric with zeros on the diagonal. The off-­diagonal
elements are also zero except that if there is direct c­oupling
between waveguides i and j, then Ψk(i,j) = Ψk(j,i) = θi,j. We call Ψk
the coupling-angle matrix of section k of the coupled waveguide
array.
Using Equation 3.179, we can express the total coupling matrix
M of the CMR network as

M = M 4 M3 M2 M1 = e jΨ4 e jΨ3 e jΨ2 e jΨ1 . (3.181)

Since the matrices M k do not commute in general, we use the Baker–


Campbell–Hausdorff formula to write the product of the matrix
exponentials in Equation 3.181 as

M = exp  j(Ψ1 + Ψ 2 + Ψ 3 + Ψ 4 ) + jX  = exp( jΨ + jX ). (3.182)

In the above equation, the matrix sum Ψ = Ψ1 + Ψ2 + Ψ3 + Ψ4


is a symmetric matrix with zeros on the diagonal. Its off-diago-
nal ­elements are given by the coupling angle between microrings

* A circular matrix M is a matrix that can be expressed in the form M = ejψ, where
ψ is a real matrix.
Coupled Microring Optical Filters 155

i and j, Ψ(i,j) = Ψ(i,j) = θi,j. The matrix X denotes the sum of all nested
commutators,

X= ∑ P (Ψ , Ψ , Ψ
m
m 1 2 3 , Ψ 4 ), (3.183)

where Pm represents polynomials of the commutators of the matri-


ces Ψk. Equation 3.182 shows that, in general, a 2D CMR network
is characterized by a coupling-angle matrix Ψ plus a commutation
matrix X. The matrix Ψ accounts for the direct couplings between
adjacent microring resonators and has the same form as the energy
coupling matrix M in the energy coupling formulation, except that
its (i, j) element is given by the coupling angle θi,j instead of the
energy coupling coefficient μi,j. As in the case of the energy coupling
matrix, the direct coupling-angle matrix Ψ explicitly reflects the 2D
coupling topology of the CMR network (i.e., the coupling topology
can be deduced from Ψ).
The commutation matrix X in Equation 3.182 accounts for the
effective couplings arising from all the indirect coupling paths
between two microrings that may or may not be neighbors. It is
in general a full but symmetric matrix, with alternating diagonal
bands of real and pure imaginary elements. In the limit of weak cou-
pling, the indirect coupling matrix can be neglected (X ≈ 0)* so that
M can be approximated by just the direct coupling term, M ≈ ejΨ.
Indeed, it can be shown that this is the approximation made in the
energy coupling formulation of 2D CMRs (Tsay and Van 2011b). We
can thus conclude that the energy coupling formulation does not
take into account the effect of the indirect couplings. For strongly
coupled CMR networks, the indirect coupling term can give rise to
resonance features that are not observed in weakly coupled systems
(Tsay and Van 2011b).
Due to the noncommutative nature of the coupling matrices
M k, a method for exactly synthesizing CMR filters of general 2D
­coupling topology in the z-domain has not been found. An approxi-
mate method that assumes that the matrices M k are commutative is

* The commutation matrix X also vanishes for microring coupling topologies


whose matrices M k do commute. One such special case is a quadruplet with
identical coupling coefficients κi,j = κ. For these structures, all the indirect cou-
pling paths cancel themselves out and the coupling matrix is given by only the
direct coupling terms.
156 Optical Microring Resonators

given in Tsay and Van (2012a). On the other hand, exact synthesis
can be achieved if one assumes a priori the coupling topology of
the CMR network. This is the approach adopted in Tsay and Van
(2011a), where an exact method for synthesizing CMR filters in
the canonical 2 × m coupling topology is developed based on the
­network order reduction approach.
In addition to their applications in realizing advanced optical
filters, 2D CMR networks can also be used to construct photonic
analogs of many 2D electronic systems, allowing many interesting
properties of these systems to be studied and explored in the optical
domain. For example, the clockwise and counterclockwise modes
in a CMR square lattice exhibit properties similar to the cyclotron
orbits of electrons in an atomic lattice subject to a magnetic field. In
particular, it has been shown that, with proper choice of the cou-
pling phases between adjacent microrings, the CMR lattice behaves
like a 2D atomic lattice subject to a uniform perpendicular mag-
netic field to the lattice (Hafezi et al. 2011, 2013, Liang and Chong
2013). In a finite CMR lattice, topologically protected edge states can
emerge which are immune to back scattering caused by imperfec-
tions in the lattice. It has been suggested that such a CMR lattice can
be used to realize photonic integrated devices such as optical delay
lines that are robust to fabrication variations (Hafezi et al. 2011).

3.8 Summary
In this chapter we developed general techniques for analyzing and
designing coupled microring optical filters. Several common cou-
pling configurations in 1D and 2D were considered and the type
of filter transfer functions realizable by each configuration was
derived. Given a set of filter specifications, the most suitable choice
of the microring filter to use is typically determined by the number
of resonators required to synthesize the desired spectral response,
the complexity of the resulting device configuration, and the sen-
sitivity of the design to parameter variations. Currently, one of the
biggest challenges to the practical realization of high-order micror-
ing filters is the issue of resonance mismatches caused by fabrica-
tion imperfections. As advances in fabrication techniques allow for
better control of device dimensions and more robust methods are
developed for post-fabrication fine-tuning of the microring reso-
nances, it is expected that increasingly more complex microring
Coupled Microring Optical Filters 157

device architectures can be realized for advanced integrated optics


filter applications.

References
Antoniou, A. 1993. Digital Filters: Analysis, Design, and Applications.
New York: McGraw-Hill.
Atia, A. E., Williams, A. E., Newcomb, R. W. 1974. Narrow-band
multiple-coupled cavity synthesis. IEEE Trans. Circuits Syst.
CAS-21(5): 649–655.
Bachman, D., Tsay, A., Van, V. 2015. Negative coupling and cou-
pling phase dispersion in a silicon quadrupole micro-racetrack
­resonator. Opt. Express 23(15): 20089–20095.
Barwicz, T., Popovic, M. A., Rakich, P. T., Watts, M. R., Haus, H. A.,
Ippen, E. P., Smith, H. I. 2004. Microring-resonator-based add-
drop filters in SiN: Fabrication and analysis. Opt. Express 12(7):
1437–1442.
Cameron, R. J. 1999. General coupling matrix synthesis methods for
Chebyshev filtering functions. IEEE Trans. Microwave Theory
Tech. 47: 433–442.
Cameron, R. J. 2003. Advanced coupling matrix synthesis techniques
for microwave filters. IEEE Trans. Microwave Theory Tech. 51: 1–10.
Chremmos, I., Uzunoglu, N. 2008. Modes of the infinite square lat-
tice of coupled microring resonators. J. Opt. Soc. Am. A 25(12):
3043–3050.
Cooper, M. L., Gupta, G., Green, W. M., Assefa, S., Xia, F., Vlasov,
Y. A., Mookherjea, S. 2010. 235-Ring coupled-resonator opti-
cal waveguides. In Conference on Lasers and Electro-Optics,
Optical Society of America, San Jose, CA, paper CTuHH3.
Grover, R., Van, V., Ibrahim, T. A., Absil, P. P., Calhoun, L. C., Johnson,
F. G., Hryniewicz, J. V., Ho, P. T. 2002. Parallel-cascaded semi-
conductor microring resonators for high-order and wide-FSR
filters. J. Lightwave Technol. 20(5): 900–905.
Ellis, M. G. 1994. Electronic Filter Analysis and Synthesis. Norwood,
MA: Artech House.
158 Optical Microring Resonators

Hafezi, M., Demler, E. A., Lukin, M. D., Taylor, J. M. 2011. Robust opti-
cal delay lines with topological protection. Nat. Phys. 7: 907–912.
Hafezi, M., Mittal, S., Fan, J., Migdall, A., Taylor, J. M. 2013. Imaging
topological edge states in silicon photonics. Nat. Photonics 7:
1001–1005.
Haus, H. A. 1984. Waves and Fields in Optoelectronics. Englewood
Cliffs, NJ: Prentice-Hall.
Heebner, J. E., Chak, P., Pereira, S., Sipe, J. E., Boyd, R. W. 2004.
Distributed and localized feedback in microresonator
sequences for linear and nonlinear optics. J. Opt. Soc. Am.
B 21(10): 1818–1832.
Hill, K. O., Meltz, G. 1997. Fiber Bragg grating technology funda-
mentals and overview. J. Lightwave Technol. 15(8): 1263–1276.
Hryniewicz, J. V., Absil, P. P., Little, B. E., Wilson, R. A., Ho, P.-T. 2000.
Higher order filter response in coupled microring resonators.
IEEE Photonics Technol. Lett. 12(3): 320–322.
Jinguji, K., Oguma, M. 2000. Optical half-band filters. J. Lightwave
Technol. 18(2): 252–259.
Lam, H. Y.-F. 1979. Analog and Digital Filters: Design and Realization.
Englewood Cliffs, NJ: Prentice-Hall.
Liang, G. Q., Chong, Y. D. 2013. Optical resonator analog of a two-
dimensional topological insulator. Phys. Rev. Lett. 110: 203904.
Liew, H. L., Van, V. 2008. Exact realization of optical transfer func-
tions with symmetric transmission zeros using the double-
microring ladder architecture. J. Lightwave Technol. 26: 2323–2331.
Little, B. E., Chu, S. T., Haus, H. A., Foresi, J., Laine, J.-P. 1997. Microring
resonator channel dropping filters. J. Lightwave Technol. 15(6):
998–1005.
Little, B. E., Chu, S. T., Hryniewicz, J. V., Absil, P. P. 2000. Filter syn-
thesis for periodically coupled microring resonators. Opt. Lett.
25(5): 344–346.
Little, B. E., Chu, S. T., Absil, P. P., Hryniewicz, J. V., Johnson, F. G.,
Seiferth, F., Gill, D., Van, V., King, O., Trakalo, M. 2004. Very
high order microring resonator filters for WDM applications.
IEEE Photonics Technol. Lett. 16(10): 2263–2265.
Coupled Microring Optical Filters 159

Madsen, C. K. 1998. Efficient architectures for exactly realizing


optical filters with optimum bandpass design. IEEE Photonics
Technol. Lett. 10: 1136–1138.
Madsen, C. K., Lenz, G. 1998. Optical all-pass filters for phase
response design with applications for dispersion compensa-
tion. IEEE Photonics Technol. Lett. 10(7): 994–996.
Madsen, C. K., Lenz, G., Bruce, A. J., Cappuzzo, M. A., Gomez,
L. T., Scotti, R. E. 1999. Integrated all-pass filters for tunable
dispersion and dispersion slope compensation. IEEE Photonics
Technol. Lett. 11(12): 1623–1625.
Madsen, C. K., Zhao, J. H. 1999. Optical Filter Design and Analysis:
A Signal Processing Approach. New York: John Wiley &
Sons Inc.
Masilamani, A. P., Van, V. 2012. Design and realization of a two-
stage microring ladder filter in silicon-on-insulator. Opt. Express
20(22): 24708–24713.
Martinez, H.G., Parks, T.W. 1978. Design of recursive digital fil-
ters with optimum magnitude and attenuation poles on the
unit circle. IEEE Trans. Acoust. Speech Signal Process. 26(2):
150–156.
Melloni, A., Morichetti, F., Martinelli, M. 2003. Linear and nonlin-
ear pulse propagation in coupled resonator slow-wave optical
structures. Opt. Quantum Electron. 35: 365–379.
Orta, R., Savi, P., Tascone, R., Trinchero, D. 1995. Synthesis of
multiple-ring-resonator filters for optical systems. IEEE
­
Photonics Technol. Lett. 7(12): 1447–1449.
Poon, J. K. S., Scheuer, J., Mookherjea, S., Paloczi, G. T., Huang, Y.,
Yariv, A. 2003. Matrix analysis of microring coupled-resonator
optical waveguides. Opt. Express 12(1): 90–103.
Poon, J. K. S., Scheuer, J., Xu, Y., Yariv, A. 2004. Designing coupled-
resonator optical waveguide delay lines. J. Opt. Soc. Am. B 21(9):
1665–1673.
Popovic, M. A. 2007. Sharply-defined optical filters and dispersion-
less delay lines based on loop-coupled resonators and ‘nega-
tive’ coupling. In Conference on Lasers and Electro-Optics, Optical
Society of America, Baltimore, MD, paper CThP6.
160 Optical Microring Resonators

Popovic, M. A., Barwicz, T., Rakich, P. T., Dahlem, M. S., Holzwarth,


C. W., Gan, F., Socci, L. et al. 2008. Experimental demonstra-
tion of loop-coupled microring resonators for optimally sharp
optical filters. In Conference on Lasers and Electro-Optics, Optical
Society of America, San Jose, CA, paper CTuNND.
Prabhu, A. M., Liew, H. L., Van, V. 2008a. Experimental determi-
nation of coupled-microring filter parameters via pole-zero
extraction. Opt. Express 16(9): 14588–14596.
Prabhu, A. M., Liew, H. L., Van, V. 2008b. Generalized parallel-­
cascaded microring networks for spectral engineering applica-
tions. J. Opt. Soc. Am. B 25: 1505–1514.
Prabhu, A. M., Van, V. 2008. Predistortion techniques for synthe-
sizing coupled microring filters with loss. Opt. Commun. 281:
2760–2767.
Tsay, A., Van, V. 2011a. A method for exact synthesis of 2xN
­coupled microring resonator networks. IEEE Photonics Technol.
Lett. 23(23): 1778–1780.
Tsay, A., Van, V. 2011b. Analytic theory of strongly-coupled micror-
ing resonators. IEEE J. Quantum Electron. 47(7): 997–1005.
Tsay, A., Van, V. 2012a. Field coupling method for the direct syn-
thesis of 2-D microring resonator networks. IEEE J. Quantum
Electron. 48(10): 1314–1321.
Tsay, A., Van, V. 2012b. Analysis of coupled microring resona-
tors in sheared lattices. IEEE Photonics Technol. Lett. 24(18):
1625–1627.
Van, V., Little, B. E., Chu, S. T., Hryniewicz, J. V. 2004. Micro-ring
resonator filters. In The 17th Annual Meeting of the IEEE Lasers
and Electro-Optics Society, Puerto Rico, Vol. 2, pp. 571–572.
Van, V. 2006. Circuit-based method for synthesizing serially
­coupled microring filters. J. Lightwave Technol. 24(7): 2912–2919.
Van, V. 2007. Synthesis of elliptic optical filters using mutually
coupled microring resonators. J. Lightwave Technol. 25(2):
­
584–590.
Van, V. 2009. Canonic design of parallel cascades of symmetric two-
port microring networks. J. Lightwave Technol. 27(2): 4870–4877.
Coupled Microring Optical Filters 161

Xia, F., Rooks, M., Sekaric, L., Vlasov, Y. 2007a. Ultra-compact high
order ring resonator filters using submicron silicon pho-
tonic wires for on-chip optical interconnects. Opt. Express 15:
11934–11941.
Xia, F., Sekaric, L., Vlasov, Y. 2007b. Ultracompact optical buffers on
a silicon chip. Nat. Photonics 1(1): 65–71.
Yariv, A., Xu, Y., Lee, R. K., Scherer, A. 1999. Coupled resonator opti-
cal waveguide: A proposal and analysis. Opt. Lett. 24: 711–713.
Yeh, P., Yariv, A., Hong, C.-S. 1977. Electromagnetic propagation
in periodic stratified media. I. General theory. J. Opt. Soc. Am.
67(4): 423–438.
Chapter 4

Nonlinear Optics Applications


of Microring Resonators

This chapter provides an overview of important nonlinear optical


processes in microring resonators and their applications. In a high-Q
microring resonator, the large buildup of light intensity at a resonance
can lead to strongly enhanced nonlinear o ­ ptical effects. These effects
have been exploited for a wide range of applications, including all-
optical switching, photonic logic operations, ­wavelength conversion,
and parametric amplification. In this chapter we will focus mainly on
intensity-dependent nonlinear effects, such as those ­arising from Kerr
nonlinearity in dielectric and semiconductor materials, although the
models and techniques developed in the chapter can also be applied to
microring devices possessing second-order nonlinearity. In addition,
in semiconductors such as silicon, significant two-photon absorption
(TPA) may also occur, which results in a large concentration of free
carriers being ­generated in the device. The effects of free carrier dis-
persion (FCD) and free ­carrier absorption (FCA) on the response of
the microring device can be significant. We will also examine these
free carrier-induced nonlinear effects in this chapter.
We will begin in Section 4.1 with a brief review of optical non-
linearity in materials and the formalisms for treating nonlinear
wave propagation in optical waveguides. Section 4.2 will examine
the effects of self-phase modulation (SPM) in a microring resonator,
which at high powers can lead to the phenomena of bistability and
self-pulsation. The application of microring resonators for all-optical
switching will be discussed in Section 4.3. In Section 4.4, we will study
the parametric processes of wavelength conversion and optical ampli-
fication based on four-wave mixing (FWM) in a microring resonator.

4.1 Nonlinearity in Optical Waveguides


4.1.1 Intensity-dependent nonlinearity
At high optical intensities, the induced polarization in a mate-
rial exhibits a nonlinear dependence on the electric field. In the

163
164 Optical Microring Resonators

frequency domain, this nonlinear dependence can be expressed in


terms of the power series

P(r , ω) = ε 0 χ(1) (ω)E(r , ω) + ε 0 χ( 2) (ω)E2 (r , ω) + ε 0 χ( 3 ) (ω)E3 (r , ω) + ,,


(4.1)

where E(r, ω) and P(r, ω) are the Fourier transforms of the electric
field E(r , t) and polarization P(r , t), and χ(n) is the nth-order suscepti-
bility. The above expression assumes that the electric field is linearly
polarized and the medium is isotropic. In the more general case, the
susceptibilities must be expressed as tensors. Specifically, the nth-
order polarization vector P(n)(r, ω) is related to the applied electric
field E(r, ω) via a susceptibility tensor of rank n + 1. For example, the
third-order polarization is given by

P( 3 ) (r , ω) = ε 0 χ ( 3 ) (ω) E(r , ω)E(r , ω)E(r , ω), (4.2)

where χ ( 3 ) is a fourth-rank tensor and ⋮ denotes the tensor product.


For materials with inversion symmetry such as amorphous solids
(e.g., SiO2, SiN) and semiconductors with centrosymmetric crystal-
line structures (e.g., Si, Ge), all the even-order terms in Equation
4.1 vanish, leaving the third-order polarization as the dominant
source of nonlinearity. On the other hand, in non-centrosymmetric
semiconductors such as GaAs and InP, both the second-order and
third-order nonlinearities are present. However, depending on the
application, one nonlinear process typically dominates over the
other in the material, so we can neglect the latter process. In this
chapter we will deal mainly with nonlinear processes that involve
only the third-order susceptibility.
A general definition of the third-order susceptibility assumes
the medium is illuminated by three waves at frequencies ωm, ωn,
and ωo. According to Equation 4.2, these waves induce a nonlin-
ear polarization at the frequency ωonm = ωo + ωn + ωm whose ith
­component is given by (Boyd 2008)
3 3 3
Pi( 3 ) (ω onm ) = ε 0 D ∑∑∑χ
j =1 k =1 l =1
1
(3)
ijkl (ω onm ; ω o , ω n , ω m )El (ω o )Ek (ω n )Ej (ω m ),

(4.3)

where {i, j, k, l} represent the x, y, and z components of the fields.


The factor D is called the degenerate factor and is equal to the
Nonlinear Optics Applications of Microring Resonators 165

number of distinct permutations of the frequencies ωm, ωn, and


­
ωo which add up to ωonm. Two important processes arise from the
­nonlinear interactions of these waves: self-phase modulation (SPM)
and cross-phase modulation (XPM). SPM refers to the nonlinear
polarization induced by a wave at f­requency ω1 on itself. This pro-
cess occurs when the ­frequencies {ωm, ωn, ωo} are equal to {ω1, ω1, −ω1},
{ω1, −ω1, ω1}, or {−ω1, ω1, ω1}, all of which give ωonm = ω1. The degener-
ate factor D is thus equal to 3 for SPM. In XPM, the nonlinear polar-
ization ­experienced by a wave at frequency ω1 is caused by another
wave at ω2. There are six ­combinations of the frequencies ω1 and
ω2 which give rise to XPM at ωonm = ω1, an example being {ωm, ωn,
ωo} = {ω1, ω2, −ω2}, so the ­degenerate factor D is equal to 6 in this case.
For isotropic materials, the susceptibility tensor χ(3) is consider-
ably simplified because of symmetry considerations. In particular,
if we assume that all the waves have the same linear polarization,
then the SPM and XPM polarizations are also parallel to the fields
and are given simply by
SPM
PNL (ω1 ) = 3ε 0 χ( 3 ) |E(ω1 )|2 E(ω1 ), (4.4)

XPM
PNL (ω1 ) = 6ε 0 χ( 3 ) |E(ω 2 )|2 E(ω1 ), (4.5)

where χ(3) is the χ(xxxx


3)
component of the susceptibility tensor.
Note that the above expressions indicate that the effect of XPM is
twice as large as that of SPM. The total polarization in the medium
is the sum of the linear and nonlinear polarizations,

P = PL + PNL = ε 0 [ χ(1) + Dχ( 3 ) |E|2 ] E, (4.6)

where D = 3 for SPM and D = 6 for XPM. From the total polarization
we can define the total relative permittivity of the medium in the
presence of third-order nonlinearity as
ε r = ε r,L + ε r,NL = 1 + χ(1) + Dχ( 3 ) |E|2 . (4.7)

In the above expression, we identify εr,L = 1 + χ(1) as the linear rel-


ative permittivity and εr,NL = Dχ(3)|E|2 as the nonlinear relative
permittivity.
The third-order nonlinearity is often expressed in terms of an
intensity-dependent refractive index defined as

n = n0 + n2 I , (4.8)
166 Optical Microring Resonators

where n0 and n2 are the linear and nonlinear refractive indices,


respectively, and I is the time-averaged optical intensity. For a plane
wave with the electric field expressed as

E(t) = E(ω)e jωt + c.c., (4.9)

the optical intensity is

I = n0 ε 0 c〈E 2 (t)〉 = 2n0 ε 0 c|E(ω)|2 . (4.10)

By equating n2 to the total relative permittivity in Equation 4.7,


we get

n2 = (n0 + n2 I )2 = 1 + χ(1) + Dχ( 3 ) |E|2 . (4.11)

Neglecting the second-order term in n2, we obtain n0 = 1 + χ(1) for


the linear refractive index and

Dχ( 3 )
n2 = (4.12)
4n02 ε 0 c

for the nonlinear index.


An alternative definition of the intensity-dependent refractive
index is by means of the expression

n = n0 + n2 〈E 2 (t)〉. (4.13)

By equating n2 〈E 2 (t)〉 = n2 I and making use of Equation 4.10,


we obtain the following relationship between n2 and n2:

Dχ( 3 )
n2 = n0 ε 0 cn2 = . (4.14)
4n0

Table 4.1 lists the values of the third-order susceptibility χ(3) and the
SPM nonlinear index n2 near the 1.55 μm wavelength for some com-
mon materials of interest in integrated optics.

4.1.2 Free carrier-induced nonlinear effects


In a semiconductor material whose energy bandgap is less than
twice the photon energy (Eg < 2ħω), TPA can occur which results in
the generation of free electrons and holes. These electrons and holes
Nonlinear Optics Applications of Microring Resonators 167

Table 4.1 Third-Order Susceptibility and Nonlinear Refractive


Index Near λ = 1.55 μm of Some Common Integrated
Optics Waveguide Materials
Material n0 χ(3) (m2/V2) n2 (cm2/W) References
SiO2 (fused 1.45 1.9 × 10 −22 2.6 × 10 −16 Eggleton et al. (2008)
silica)
SiN 2.0 3.4 × 10−21 2.4 × 10−15 Ikeda et al. (2008)
Si 3.45 1.9 × 10−19 4.5 × 10−14 Dinu et al. (2003)
AlGaAs 3.14 1.3 × 10−19 3.6 × 10−14 Islam et al. (1992)
GaAs 3.47 6.8 × 10−19 1.6 × 10−13 Islam et al. (1992)

modify the medium’s optical properties through the FCA and FCD
effects, leading to a nonlinear change in the absorption and refrac-
tive index of the medium. For example, at the telecommunication
wavelengths, considerable TPA can occur in silicon at moderate to
high power levels, which gives rise to nonnegligible free carrier-
induced nonlinear effects.
The effects of free carriers on the optical polarization can be
accounted for by defining a nonlinear susceptibility due to free
­carriers, χfc, as

 ∆α fc 
χfc = 2n0  ∆nfc − j , (4.15)
 2k 

where k = ω/c, n0 is the linear refractive index, and Δnfc and Δαfc
are the changes in the refractive index and absorption, respectively,
induced by the free carriers. These changes are related to the gener-
ated electron and hole densities, Ne and Nh, in the medium accord-
ing to the relations

∆nfc = σ (re ) N e + σ (rh ) N h , (4.16)

∆α fc = σ (ae ) N e + σ (ah ) N h , (4.17)

where σ (re , h ) and σ (ae , h ) are the refraction volume and absorption
cross section, and the superscripts (e) and (h) indicate the contribu-
tions from electrons and holes, respectively. Since TPA generates an
equal number of electrons and holes, we can set Ne = Nh = Nfc in the
above relations to get

∆nfc = σ (re ) + σ (rh )  N fc = σ r N fc , (4.18)


168 Optical Microring Resonators

∆α fc = σ (ae ) + σ (ah )  N fc = σ a N fc , (4.19)

where σr and σa denote the total free carrier (FC) refraction volume
and absorption cross section, respectively. For Si at the 1.55 μm
wavelength, the following empirical formulas are often used (Soref
and Bennett 1987):

∆nfc = −(8.8 × 10 −22 N e + 8.5 × 10 −18 N h0.8 ) ≈ σ r N fc , (4.20)

∆α fc = 8.5 × 10 −18 N e + 6.0 × 10 −18 N h = σ a N fc , (4.21)

where Ne and Nh are in units of cm−3 and Δαfc is in units of cm−1.


The value of the total absorption cross section σa in Equation 4.21
is 1.45 × 10−17 cm2. The total refraction volume σr in Equation 4.20
may be determined by considering that at electron densities around
1016−1017 cm−3, for which FC effects in silicon become nonnegligible,
the contribution of holes to Δnfc is about 5 times that of electrons.
One may thus use an effective value of σr = −5.3 × 10−21 cm3 for the
refraction volume at these carrier densities (Lin et al. 2007).
The free carrier densities generated by TPA are directly propor-
tional to the squared intensity of the absorbed light. Due to diffu-
sion and recombination, the generated carriers have a finite lifetime,
so the induced nonlinear effects in the medium also exhibit a finite
response time. In the absence of an external applied field, the
­spatial and temporal dependence of the FC density, Nfc(r, t), gener-
ated by TPA is governed by the diffusion equation

∂N fc N fc α
+ − D∇ 2 N fc = 2 I 2 , (4.22)
∂t τ rec 2ω

where I is the optical intensity, α2 the TPA coefficient, τrec the carrier
lifetime due to recombination, and D is the diffusion constant. At
the 1.55 μm wavelength, the TPA coefficient of Si is in the range of
0.5−0.8 cm/GW (Dinu et al. 2003, Lin et al. 2007). The recombina-
tion lifetime τrec in bulk Si is typically in the range of a few micro-
seconds (Dimitropoulos et al. 2005). If we define a diffusion time
constant τdiff such that D∇2 Nfc = −Nfc/τdiff, then Equation 4.22 can be
expressed as

∂N fc N fc α
+ = 2 I2 , (4.23)
∂t τ fc 2ω
Nonlinear Optics Applications of Microring Resonators 169

where the effective FC lifetime τfc is given by


1 1 1
= + . (4.24)
τ fc τ rec τ diff

In a silicon waveguide, the effective FC carrier lifetime τfc depends


on the waveguide geometry, dimensions, and the defect level in the
material and is typically in the order of a nanosecond (Dimitropoulos
et al. 2005, Turner-Foster et al. 2010).

4.1.3 Wave propagation in a nonlinear optical waveguide


The formalisms for analyzing nonlinear guided-wave propagation
have been extensively developed for optical fibers and waveguides
(Marcuse 1991, Agrawal 2013). Here we use a similar approach to
Agrawal (2013) and Lin et al. (2007) to derive the nonlinear wave
equation for the propagation of a pulse in an integrated optical
waveguide with an instantaneous intensity-dependent refractive
index. Nonlinear effects due to TPA and free carriers will be con-
sidered in Section 4.1.4.
We consider a waveguide oriented along the z direction with a
cross-sectional linear index distribution n0(x, y). For simplicity we
assume that both the core and cladding materials are isotropic. The
core material exhibits an instantaneous intensity-dependent refrac-
tive index of the form given by Equation 4.8. We will neglect any
nonlinear effect in the cladding region, either because the material
nonlinearity is weak or because the waveguide is strongly confined
so that the field in the cladding is small. The starting point in our
derivation is the nonlinear wave equation*

1 ∂2E ∂ 2 PL ∂ 2 PNL
∇2E − = µ 0 + µ 0 , (4.25)
c 2 ∂t 2 ∂t 2 ∂t 2

where E(r , t) represents a linearly polarized electric field, and PL (r , t)


and PNL (r , t) are the linear and nonlinear polarizations, respectively.
Taking the Fourier transform of Equation 4.25, we get

∇ 2 E(r , ω) + k 2 E(r , ω) = −µ 0 ω 2 PL (r , ω) − µ 0 ω 2 PNL (r , ω), (4.26)

* Equation 4.25 neglects the vectorial nature of the waveguide mode, which could
become important for high-index contrast waveguides in certain applications.
A full-vectorial treatment must include the term −∇(∇ ⋅ E ) on the left-hand side
of Equation 4.25 to take into account polarization coupling.
170 Optical Microring Resonators

where k = ω/c and E(r, ω) is the Fourier transform of the electric field
defined as

E(r , ω) =
∫ E(r, t)e
−∞
− jωt
dt. (4.27)

The Fourier transforms PL(r, ω) and PNL(r, ω) of the linear and non-
linear polarizations are similarly defined and can be related to the
electric field by

PL (r , ω) = ε 0 χ(1) ( x , y , ω)E(r , ω), (4.28)

PNL (r , ω) = ε 0 ε r,NL ( x , y , ω)E(r , ω). (4.29)

In the above equations, χ(1) and εr,NL are the linear susceptibility and
nonlinear relative permittivity, respectively, which are both func-
tions of (x, y) to account for the spatial index distribution of the
waveguide cross section. Upon substituting Equations 4.28 and 4.29
into Equation 4.26, we obtain
∇ 2 E(r , ω) + n02 k 2 E(r , ω) = − ε r,NL k 2 E(r , ω), (4.30)

where n02 = 1 + χ(1) . Equation 4.30 is the nonlinear wave equation for
the electric field in the frequency domain. The term on the right-
hand side describes the effect of the nonlinear permittivity on light
propagation in the waveguide.
We now solve Equation 4.30 for the case of a pulse signal
­modulating a carrier frequency ω0. For a pulse propagating in the
z direction with a bandwidth centered around frequency ω0, we can
express the electric field in the waveguide as
 ( z , t)e j(ω0t −β0 z ) ,
E(r , t) = φ( x , y )A (4.31)

where ϕ(x,y) is the waveguide mode and Ã(z, t) is the slowly varying
pulse envelope. The Fourier transform of Equation 4.31 is

E(r , ω) = φ( x , y )A( z , ω − ω 0 )e − jβ0 z , (4.32)

where

∫ A (z, t)e
− j( ω − ω 0 )t
A( z , ω − ω 0 ) = dt. (4.33)
−∞
Nonlinear Optics Applications of Microring Resonators 171

The waveguide mode satisfies the eigenvalue equation

∇ T2 + n02 ( x , y , ω)k 2  φ( x , y ) = β 2 (ω)φ( x , y ), (4.34)

where ∇ T2 is the transverse Laplacian operator and β(ω) is the linear


propagation constant. In Equation 4.31, we define the propagation
constant β0 to be the value of β(ω) at the center frequency, β0 = β(ω0).
By substituting Equation 4.32 into the wave equation (4.30) and
­performing separation of variables, we obtain

∇ T2 φ + (n02 + ε r,NL )k 2 φ = β NL
2
φ, (4.35)

∂A
j 2β0 − (β 2NL − β02 )A = 0, (4.36)
∂z

where we have made use of the slowly varying envelope approx-


imation ∂2 A/∂z2 ≪ jβ0∂A/∂z in Equation 4.36. From Equation 4.35
we identify the separation constant βNL as the nonlinear propaga-
tion constant of the waveguide in the presence of the nonlinear
relative permittivity εr,NL. The value of βNL can be determined by
expressing it as a perturbation of the linear propagation constant
β: βNL = β(ω) + ΔβNL. Making the approximation

β 2NL = (β + ∆β NL )2 ≈ β 2 + 2β∆β NL , (4.37)

and substituting the result into Equation 4.35, we get

ε r,NL k 2 φ( x , y ) = 2β∆β NL φ( x , y ). (4.38)

Upon multiplying Equation 4.38 by ϕ*(x, y) and integrating over the


transverse plane, we obtain

∆β NL =
k2 ∫εC
r,NL |φ( x , y )|2 dx dy
. (4.39)

∫|φ(x, y)| dx dy
2

Note that the integral in the numerator is taken only over the cross
section C of the core since we assume that only the waveguide core
exhibits nonlinearity.
172 Optical Microring Resonators

Next we substitute the solution for ΔβNL into Equation 4.36 to


obtain the equation for the pulse envelope A(z, ω − ω0). Using the
approximation
2
β NL − β02 = (β NL + β0 )(β NL − β0 ) ≈ 2β0 (β + ∆β NL − β0 ), (4.40)

we can simplify Equation 4.36 as follows:

∂A
+ j(β − β0 )A = − j∆β NL A. (4.41)
∂z

The above equation describes the propagation of the pulse envelope


in the frequency domain. Chromatic dispersion effects are mani-
fested through the frequency dependence of the propagation con-
stant, β(ω). To account for linear loss in the waveguide, we can add
the term α0A/2, where α0 is the linear absorption coefficient, to the
left-hand side of Equation 4.41:

∂A α
+ j[β(ω) − β0 ]A + 0 A = − j∆β NL A. (4.42)
∂z 2

For narrowband pulses, we may assume that the linear loss is


approximately constant over the frequency spectrum of the signal.
To convert the nonlinear wave equation (4.42) into the time
domain, we expand β(ω) in terms of a Taylor series around ω0 to get
(Agrawal 2013)

∂A α
+ j(∆ωβ1 + 12 ∆ω 2β 2 + )A + 0 A = − j∆β NL A, (4.43)
∂z 2

where Δω = ω − ω0 and βn is the nth-order derivative of β:


β n = (d nβ/dω n )ω =ω0. Taking the inverse Fourier transform of the
above equation and keeping the first two terms in the Taylor series,
we obtain


∂A  jβ 2 ∂ 2 A
∂A  α0
+ β1 − +  = − j∆β NL A
A , (4.44)
∂z ∂t 2 ∂t 2 2

where Ã(z, t) is the inverse Fourier transform of A(z, ω − ω0) as


defined by Equation 4.33. Equation 4.44 is the nonlinear wave equa-
tion for pulse propagation in the time domain. The terms β1 and β2
account for the effects of phase velocity dispersion (or group delay)
Nonlinear Optics Applications of Microring Resonators 173

and group velocity dispersion (GVD), respectively. Equation 4.44


is general in that the term ∆β NL can account for any type of non-
linearity in the waveguide core, including free carrier-induced
nonlinearity.
For the case where the waveguide core has Kerr nonlinearity,
the nonlinear relative permittivity is given by εr,NL = 3χ(3)|E(r, ω)|2.
Substituting this expression into Equation 4.39 and making use of
Equation 4.32 for the electric field, we obtain the nonlinear contri-
bution to the propagation constant as follows:

∆β NL =
3 χ( 3 ) k 2 ∫ |φ(x, y)| dx dy |A(z, ω − ω )| .
C
4

0
2
(4.45)

∫|φ(x, y)| dx dy
2

Taking the inverse Fourier transform of the above expression, we


get


∆β NL =
3 χ( 3 ) k C ∫
|φ( x , y )|4 dx dy
 ( z , t)|2 ,
|A (4.46)

2neff |φ( x , y )|2 dx dy

where neff = cβ0/ω0 is the effective index of the waveguide and we


have also made the approximation β ≈ neffk.
It is convenient to normalize the pulse amplitude à so that |Ã|2
gives the power in the waveguide. Since the time-averaged power
in the waveguide is given by

∫ ∫
 ( z , t)|2 |φ( x , y )|2 dx dy ,
P( z , ω) = neff ε 0 c 〈E 2 (r , t)〉dx dy = 2neff ε 0 c|A

(4.47)
we normalize Equation 4.44 by making the substitution

A
→
A .
( )
1/2


2neff ε 0 c |φ( x , y )|2 dx dy

The resulting equation is


∂A  jβ 2 ∂ 2 A
∂A  α0
+ β1 − +  = − j γ |A
A  |2 A
, (4.48)
∂z ∂t 2 ∂t 2 2
174 Optical Microring Resonators

where the nonlinear coefficient γ is given by

3 χ( 3 ) ω
γ= 2
, (4.49)
4neff ε 0 c 2 Aeff

and Aeff is the effective mode area of the waveguide defined as

(∫|φ(x, y)| dx dy ) .
2
2

Aeff = (4.50)
∫ C
|φ( x , y )|4 dx dy

From Equation 4.12 we obtain the relationship between χ(3) and the
nonlinear index n2 in the core material as χ( 3 ) = 4nc2 n2 ε 0 c/3, so the
nonlinear coefficient γ is also given by*

2
 n  nω (4.51)
γ = c  2 .
 neff  cAeff

For high-index contrast waveguides, the core index can be substan-


tially larger than the effective index, leading to a factor (nc/neff)2
enhancement in the nonlinear coefficient γ compared to that in low-
index contrast waveguides. Additional enhancement also comes
from the reduced effective mode area Aeff of strongly confined
waveguides compared to those with weak modal confinement.
Analytical solution of the nonlinear wave equation (4.48) can
be obtained for some simple cases. In particular, for narrowband
pulses, we can neglect the effects of phase velocity dispersion and
GVD and solve the simplified equation

 α0
∂A
+  = − j γ |A
A  |2 A
. (4.52)
∂z 2

* In most of the literature, the nonlinear coefficient is given as γ = n2ω/cAeff, with-


out the factor (nc/neff)2. This is a good approximation for low-index contrast
waveguides. However, for high-index contrast waveguides where the effective
index is significantly smaller than the core index, analytical solutions of the non-
linear wave equation in Equation 4.48 are in better agreement with numerical
simulations based on the exact equation in (4.25) if the factor (nc/neff)2 is included
in the nonlinear coefficient.
Nonlinear Optics Applications of Microring Resonators 175

Writing Ã(z, t) = U(z, t)e jφ(z,t), where U(z,t) = |Ã(z, t)|, we substitute
it into Equation 4.52 and separate the real and imaginary parts to
get

∂U α 0
+ U = 0, (4.53)
∂z 2

∂ϕ
= − γU 2 ( z , t). (4.54)
∂z

From Equation 4.53 we obtain the solution for the pulse amplitude,

U( z , t) = U(0, t)e − α0 z/2 . (4.55)

Substituting the above result into Equation 4.54 and solving for the
nonlinear phase φ, we get

ϕ( z , t) = ϕ(0, t) − γU 2 (0, t)zeff , (4.56)

where zeff is the effective propagation distance given by

1 − e −α0 z
zeff = . (4.57)
α0

The solution for the pulse envelope can thus be expressed as

 ( z , t) = A
A  (0, t)e − α0 z/2 e jϕ ( z ,t ) , (4.58)

with the nonlinear phase given by

 (0, t)|2 zeff .


ϕ( z , t) = −γ |A (4.59)

Equation 4.58 shows that as the pulse propagates in the nonlinear


waveguide, it retains the initial temporal shape of the ­envelope
except for a decrease in the amplitude due to linear loss. On the other
hand, the signal acquires a nonlinear phase shift φ, which depends
on the initial pulse power profile |Ã(0, t)|2. In ­particular, since φ
changes with time, the pulse acquires a frequency chirp given by
δω = ∂φ/∂t, which leads to spectral b ­ roadening of the signal as it
propagates along the nonlinear waveguide.
176 Optical Microring Resonators

For a monochromatic wave at frequency ω0, the solution is


simply

A( z) = A0 e − α0 z/2 e jϕ ( z ) , (4.60)

where the nonlinear phase shift is φ(z) = −γ|A0|2 zeff. In a micror-


ing resonator, the acquired nonlinear phase due to SPM causes
a shift in the resonant frequency. The total nonlinear phase
accumulated over one round-trip of the microring is ϕNL =
­
−γ|A0|2 L eff, where L eff is the effective circumference of the micror-
ing defined as

1 − e −2α0 πR  a2 − 1 
Leff = = 2πR  rt . (4.61)
α0  2 ln art 

In the above expression, art = e −α0 πR is the round-trip amplitude


attenuation in the microring. To determine the shift in the reso-
nant frequency caused by the nonlinear phase, we note that, for
a change ΔnNL in the effective index of the microring waveguide,
the resonant frequency ωm of resonance mode m experiences a
shift given by ΔωNL = −(ωm/ng)ΔnNL, where ng is the group index. By
writing the nonlinear round-trip phase as ϕNL = −ΔnNL(ωm/c)2πR =
−ΔnNL(2mπ/nr), where nr is the effective index, we obtain the follow-
ing e­ xpression for the resonant frequency shift:

n  φ ω n 
∆ω NL = ω m  r  NL = − m  r  γ |A0 |2 Leff . (4.62)
 ng  2mπ 2mπ  ng 

The above equation shows that the resonant frequency is red shifted
if the nonlinear coefficient γ (or n2) is positive, which is t­ ypically the
case for Kerr nonlinearity in most integrated optics materials near
the 1.55 μm wavelength. We will study SPM effects in microring
resonators in more detail in Section 4.2.

4.1.4 Free carrier-induced nonlinear effects


in an optical waveguide
In Section 4.1.2 we gave a brief review of the nonlinear effects
­associated with free carriers generated by TPA in a semiconduc-
tor material whose bandgap is less than twice the photon energy
(Eg < 2ħω). To incorporate these effects into the nonlinear wave
Nonlinear Optics Applications of Microring Resonators 177

equation, we write the total nonlinear relative permittivity in the


waveguide core as
ε r,NL = ε r,K + ε r,fc , (4.63)

where εr,K and εr,fc are the contributions from Kerr nonlinearity
and free carriers, respectively. TPA manifests itself as an intensity-
dependent nonlinear loss, which can be accounted for by adding an
imaginary term to the Kerr coefficient, n2 − jα2/2k, where α2 is the
TPA coefficient and k = ω/c. The nonlinear relative permittivity due
to the Kerr effect is thus given by

ε r,K = 3χ( 3 ) |E(r , ω)|2 = 4 nc2 ε 0 c(n2 − jα 2 /2k )|E(r , ω)|2 , (4.64)

where nc is the linear refractive index of the waveguide core. The con-
tribution of free carriers to the nonlinear permittivity is given by the
FC susceptibility defined in Equation 4.15. Using Equations 4.18 and
4.19, we can write the nonlinear relative ­permittivity due to free
­carriers in terms of the refraction volume and absorption cross
­section as
ε r,fc = χfc = 2nc (σ r − jσ a /2k )N fc , (4.65)

where Nfc(r, t) is the FC density in the waveguide core. We can now


evaluate the contribution of free carriers to the nonlinear propaga-
tion constant of the waveguide by using Equation 4.39,

∆β fc =
k2 ∫εC
r,fc |φ( x , y )|2 dx dy
=
nc k
(σ r − jσ a /2k )N fc , (4.66)

∫ neff
2
|φ( x , y )| dx dy

where N fc ( z , t) is the spatially averaged FC density over the cross


section of the waveguide core,

N ( z , t) =
fc
∫ N (r, t)|φ(x, y)| dx dy .
C
fc
2

(4.67)
∫|φ( x , y )|2 dx dy

Combining the Kerr and FC-induced nonlinear effects, we can write


the nonlinear wave equation as

∂A  jβ 2 ∂ 2 A
∂A  α0
+ β1 − +  = − jγ K |A
A |2 A
 − jγ fc N fc A
, (4.68)
∂z ∂t 2 ∂t 2 2
178 Optical Microring Resonators

where the nonlinear Kerr and FC coefficients are given by

nc2 k  jα 2 
γK = 2  n2 − 2k  , (4.69)
neff Aeff

nc k  jσ a 
γ fc =  σr − . (4.70)
neff  2k 

The above expressions show that, in an optical waveguide,


Kerr nonlinear effects are enhanced by a factor of (nc/neff)2, whereas
FC-induced effects are enhanced by a factor of (nc/neff). Also note
that since the FC refraction volume σr is negative while the Kerr
coefficient n2 is typically positive, the nonlinear phase shifts due to
FC and Kerr effects are in opposite directions to each other.
In the absence of an external applied field, the free carri-
ers ­generated from TPA diffuse throughout the waveguide core
according to Equation 4.22. Taking the spatial average of this
­equation over the waveguide cross section, we can write the result
as (Lin et al. 2007)

∂N fc  1 1 
+ + N fc = G, (4.71)
∂t  τ rec τ diff 

where G is the spatially averaged FC generation rate and τdiff is the


diffusion time constant defined by

∫ D∇ N
C
2
fc |φ( x , y )|2 dx dy
=−
N fc
. (4.72)

|φ( x , y )| dx dy τ diff
2

The FC generation rate G can be computed from the attenuation of


the optical intensity due to TPA,

G=−
1 ∂I
=−
1 ∫ (∂I/∂z)|φ(x, y)| dx dy .
C
2

(4.73)
2ω ∂z 2ω

|φ( x , y )|2 dx dy

Since the optical intensity of the waveguide mode is


2 
I ( x , y , z) = neff ε 0 c〈E 2 (r , t)〉 = 2neff ε 0 c|φ( x , y )|| A( z , t)|2 , (4.74)
Nonlinear Optics Applications of Microring Resonators 179

we have

∂I  ∂U 
= 2neff ε 0 c|φ( x , y )|2  2U , (4.75)
∂z  ∂z 

where U = |Ã(z, t)|. The rate of attenuation of the field amplitude


due to TPA can be determined from

∂U
= Im{∆β K }U , (4.76)
∂z

where, according to Equation 4.46,

Im{∆β K } = −
nc
2
ε 0 cα 2 ∫
 |φ( x , y )|4 dx dy 
 C  2
 U . (4.77)
neff

2
 |φ( x , y )| dx dy 

Using the above results in Equation 4.73, we obtain the


­following expression for the average FC generation rate:
2

G=
nc2α 2 2 C
( 2ε 0 c) 

 |φ( x , y )|4 dx dy 
 4 (4.78)
 U .
2ω

 |φ( x , y )| dx dy 
2

Upon performing power normalization of the pulse envelope, we


finally get
2
 n  α 2U 4 (4.79)
G= c  = ζ TPAU 4 ,
 neff  2ωAeff
2

where the effective mode area Aeff is given in Equation 4.50.


The equation governing the time evolution of the average FC
­density in Equation 4.71 can now be expressed as

∂N fc N fc  ( z , t)|4 ,
+ = ζ TPA |A (4.80)
∂t τ fc

where τfc is the effective FC lifetime defined by

1 1 1
= + . (4.81)
τ fc τ rec τ diff
180 Optical Microring Resonators

In general, the effects of FCA and FCD become important only


for optical pulses with high peak powers or large pulse widths com-
pared to the carrier lifetime. To quantify the relative magnitudes
of FCA and TPA, Lin et al. (2007) considered the propagation of a
Gaussian pulse with initial power profile |A(0, t)|2 = P0 exp(−t 2 /T02 )
and determined the ratio of the maximum absorption due to
FCA (α FCA TPA
max ) to that due to TPA (α max ) . The result can be expressed as

α FCA
max n σ a Ep
ra = TPA
= c , (4.82)
α max neff 2 2ωAeff

where E p = πP0 T0 is the energy of the Gaussian pulse. The above


expression shows that ra depends only on the pulse energy. FCA
effects can be neglected if ra≪1, which is typically satisfied for pulse
energy E p < 30 pJ in a Si waveguide around the 1.55 μm wavelength.
A similar ratio can also be defined to assess the relative magni-
tude of FC-induced index change to that due to Kerr nonlinearity.
Since an important manifestation of the nonlinear index change is
spectral broadening due to frequency chirping, one can look at the
growth rate of the frequency chirp with respect to the propagating
distance,

∂(δω) ∂  ∂ϕ 
C= =  . (4.83)
∂z ∂z  ∂t 

FC
The ratio of the maximum chirp growth due to FCD (Cmax ) to that
Kerr
due to Kerr nonlinearity (Cmax ) is found to be (Lin et al. 2007)
FC
Cmax n α 2 |σ r|Ep
rC = Kerr
= c , (4.84)
Cmax neff 2 π ωn2 Aeff

which also depends only on the pulse energy. The effects of


FC-induced index change can be neglected if rC ≪ 1, which is
typically satisfied for pulse energy E p < 10 pJ in a Si waveguide
around 1.55 μm.
For the general case where both the effects of Kerr nonlinear-
ity and free carriers are nonnegligible, solution of the coupled
differential equations in Equations 4.68 and 4.80 can be obtained
using a numerical method such as the finite difference method.
When the effect of FCA is small ( γ fc′′ = Im{γ fc } ≈ 0), a semi-analytical
­solution can also be obtained if phase velocity dispersion and GVD
Nonlinear Optics Applications of Microring Resonators 181

are neglected (β1 = 0, β2 = 0). Writing the solution for the pulse enve-
lope as Ã(z, t) = U(z, t)e jφ(z,t) and substituting it into Equation 4.68,
we get

∂U α 0
+ U = − γ K′′U 3 , (4.85)
∂z 2

∂ϕ
= − γ K′ U 2 − γ fc′ N fc , (4.86)
∂z

where γ K′ and γ K′′ are the real and imaginary parts of the Kerr
nonlinear coefficient and γ fc′ is the real part of the FC coefficient.
We solve Equation 4.85 first to obtain the following expression for
the pulse power, P(z, t) = U2(z, t):

Pin (t)e − α0 z
P( z , t) = , (4.87)
1 + 2γ K′′ zeff Pin (t)

where zeff is the effective propagation distance given in Equation


4.57 and Pin(t) = P(0, t) is the initial power profile of the pulse. Next
we substitute Ã(z, t) = U(z, t)e jϕ(z,t) into Equation 4.80 to get the equa-
tion for the FC density,

dN fc N fc
+ = ζ TPA P 2 ( z , t). (4.88)
dt τ fc

The above equation can be solved to give


t

N fc ( z , t) = ζ TPA e − t/τ fc
∫ P (z, t′)e
−∞
2 t ′/τ fc
dt ′ = ζ TPAQ( z , t)e −2α0 z e − t/τfc ,

(4.89)
where
t
Pin2 (t ′)e t ′/τfc dt ′
Q( z , t) = ∫
−∞
[1 + 2γ K′′ zeff Pin (t ′)]2
. (4.90)

The integral in Equation 4.90 can be numerically computed for a


given initial pulse power profile Pin(t). Substituting Equation 4.89
for the FC density N fc into Equation 4.86, we can now solve for the
nonlinear phase shift φ. The result can be expressed as φ = φK + φfc,
182 Optical Microring Resonators

where φK and φfc are the nonlinear phase shifts due to Kerr and FC
nonlinearity, respectively:
z


ϕ K ( z , t) = − γ K′ P( z ′ , t)dz ′ ,
0
(4.91)


ϕ fc ( z , t) = − γ fc′ ζ TPA e − t/τfc Q( z ′ , t)e −2α0 z ′ dz ′.
0
(4.92)

The effects of FCA and FCD on the propagation of a Gaussian


pulse at the 1.55 μm wavelength in a Si waveguide are illustrated in
Figure 4.1. The waveguide has a silicon core (nc = 3.45) with cross-
sectional dimensions 450 × 250 nm2 surrounded by SiO2 cladding.
The effective index of the fundamental TE mode is neff = 2.5 and the
effective mode area is Aeff = 0.15 μm2. The linear propagation loss in
the waveguide is assumed to be 2.5 dB/cm. The Kerr and FC para­
meters for Si at 1.55 μm are n2 = 4.5 × 10−14 cm2/W, α2 = 0.8 cm/GW,
σr = −5.3 × 10−21 cm3, and σa = 1.45 × 10−17 cm2. We assume the carrier
lifetime τfc in the Si waveguide to be 1 ns. The Gaussian pulse has an
initial power profile given by Pin (t) = P0 exp(−t 2 /T02 ).
Figures 4.1a–c show the pulse power (P(t) = |A(z,t)|2), nonlin-
ear phase (φ(t)), and FC carrier density ( N fc (t)) in the waveguide
after a propagation distance of 1 mm. The results are obtained for
a fixed peak input power P0 = 1 W but different pulse widths T0
of 0.01 ns, 0.1 ns, and 1 ns, which correspond to pulse energies of
1.8 pJ, 18 pJ, and 180 pJ, respectively. For the 1.8 pJ pulse, the ratios
ra and rC given by Equations 4.82 and 4.84 are calculated to be
0.065 and 0.34, respectively, indicating that TPA and Kerr-induced
index change dominate over FC effects. In particular, we observe
that the nonlinear phase is negative during the leading edge of the
pulse before becoming positive as the FCD effect becomes more
prominent. For higher pulse energies, both nonlinear absorption
and index change are dominated by FC effects, with the nonlin-
ear phase and FC density exhibiting long tails due to the large
carrier lifetime. Figures 4.1d–f show the effects of increased peak
pulse power P0 for a fixed pulse width T0 of 0.1 ns. The values of P0
are 1 W, 2 W, and 5 W, corresponding to pulse energies of 0.18 nJ,
0.36 nJ, and 0.9 nJ, respectively. At high pulse energies, strong
FCA causes the pulse shape to become severely attenuated and
distorted, as evident in Figure 4.1d.
Nonlinear Optics Applications of Microring Resonators 183
(a) (d)
1 0.7
0.01 ns 1W
0.1 ns 0.6 2W
0.8 1 ns 5W
0.5

Power, P(t)/P0
Power, P(t)/P0

0.6 0.4

0.4 0.3
0.2
0.2
0.1
0 0
–3 –2 –1 0 1 2 3 –3 –2 –1 0 1 2 3
Time t/T0 Time t/T0
(b) (e)
0.14 8
0.01 ns 1W
0.12 0.1 ns 7 2W
Nonlinear phase, φ(t)/π

Nonlinear phase, φ(t)/π


1 ns 5W
0.1 6
0.08 5
0.06 4
0.04 3
0.02 2
0 1
–0.02 0
–2 0 2 4 6 8 –2 0 2 4 6 8 10
Time t/T0 Time t/T0

(c) × 1015 (f) × 1017


14 2
0.01 ns 1W
12 0.1 ns 2W
FC density, Nfc(t)(cm–3)

1 ns
FC density, Nfc(t)(cm–3)

5W
1.5
10
8
1
6
4 0.5
2
0 0
–2 0 2 4 6 8 –2 0 2 4 6 8 10
Time t/T0 Time t/T0

Figure 4.1 Power, nonlinear phase shift, and FC density in a Si waveguide


after 1 mm propagation distance for: (a)–(c) fixed peak input pulse power
P0 = 0.1 W and pulse width T0 = 0.01 ns, 0.1 ns, and 1 ns, (d)–(f) fixed input
pulse width T0 = 0.1 ns and peak input pulse power P0 = 1 W, 2 W, and 5 W.

4.1.5 FWM in a nonlinear optical waveguide


When two or more light waves at different frequencies propagate
in a nonlinear optical waveguide, a variety of processes such as
XPM, parametric amplification, and frequency conversion can
occur. In this section we consider the process of four wave mixing
184 Optical Microring Resonators

(FWM), which involves the interaction of four monochromatic


waves at frequencies ω1, ω2, ω3, and ω4 in an optical waveguide
with intensity-dependent refractive index. In a typical FWM
scheme, two waves called the pump beams at frequencies ω1
and ω2 and a signal wave at frequency ω3 are launched into the
waveguide. FWM then gives rise to an idler wave at frequency ω4.
Energy conservation requires that the ­generated frequency sat-
isfy the relation ω4 = ω1 + ω2 − ω3.
To analyze the FWM process, we assume that the wave-
guide is oriented along the z direction with linear refractive
index ­distribution over its cross section given by n0(x, y). Both the
core and ­cladding materials are isotropic, with the core material
­possessing Kerr nonlinearity. The two pump beams and the signal
wave co-propagate in the waveguide in the positive z direction and
are assumed to have the same linear polarization. Consequently,
the generated idler wave will also have the same linear polarization
and propagate in the same direction. We express the electric field of
the ith wave at frequency ωi in the waveguide as

Ei (r , t) = Ei (r )e jωit + c.c.. (4.93)

Due to nonlinear interactions of the four waves, the total nonlinear


polarization in the waveguide will contain components at various
combinations of the four frequencies. However, only those compo-
nents at frequency ωi will drive the electric field of the ith wave.
Writing the nonlinear polarization at frequency ωi as

PNL (r , t , ω i ) = PNL (r , ω i )e jωit + c.c., (4.94)

and substituting the expressions in Equations 4.93 and 4.94 into


the wave equation (4.25), we obtain the equation for the field
amplitude Ei(r),

∇ 2 Ei (r ) + n02 ki2 E(r ) = −µ 0 ω i2 PNL (r , ω i ), (4.95)

where ki = ωi/c. Next, we write the electric field Ei(r) in the wave-
guide in terms of the waveguide mode ϕi(x, y) and amplitude Ai(z),

Ei (r ) = φi ( x , y )Ai ( z)e − jβi z , (4.96)


Nonlinear Optics Applications of Microring Resonators 185

where βi = β(ωi) is the propagation constant at frequency ωi.


The transverse modal distribution ϕi(x, y) satisfies the eigenvalue
equation

∇ T2 φi ( x , y ) + n02 ki2 = βi2 φi ( x , y ), (4.97)

where ∇ T2 denotes the transverse Laplacian operator. By substitut-


ing Equation 4.96 into Equation 4.95, we get

dAi − jβi z
− j 2βi φi ( x , y ) e = −µ 0 ω i2 PNL (r , ω i ), (4.98)
dz

where we have made use of Equation 4.97 and the slowly


­vary­i ng amplitude approximation d2 Ai/dz2 ≪ jβidAi/dz. Upon
multi­ply­i ng Equation 4.98 by φ∗i ( x , y ) and integrating over the
transverse plane, we obtain the equation for the amplitude of
the ith beam,


dAi jµ 0 ω i2 PNL (r , ω i ), φi jβi z
=− e , (4.99)
dz 2βi |φi |2

where the angled brackets denote integration over the x–y plane.
In a medium with Kerr nonlinearity, the total nonlinear polar-
ization is given by

PNL (r , t) = ε 0 χ( 3 )E(r , t)E(r , t)E(r , t), (4.100)

where E = E 1 + E 2 + E 3 + E 4 is the total electric field. Writing each of


the fields E i in the form of Equation (4.93) and substituting the total
field into Equation 4.100, we obtain the following expressions for
the nonlinear polarization at each frequency ωi:

( )
PNL (r , ω1 ) = 3ε 0 χ( 3 ) |E1|2 E1 + 2 |E2|2 +|E3|2 +|E4|2 E1 + 2E2∗E3 E4  ,

(4.101)

( )
PNL (r , ω 2 ) = 3ε 0 χ( 3 ) |E2|2 E2 + 2 |E1|2 +|E3|2 +|E4|2 E2 + 2E1∗E3 E4  ,

(4.102)
186 Optical Microring Resonators

( )
PNL (r , ω 3 ) = 3ε 0 χ( 3 ) |E3 |2 E3 + 2 |E1 |2 +|E2 |2 +|E4 |2 E3 + 2E1E2 E4∗  ,

(4.103)

( )
PNL (r , ω 4 ) = 3ε 0 χ( 3 ) |E4 |2 E4 + 2 |E1 |2 +|E2 |2 +|E3 |2 E4 + 2E1E2 E3∗  .

(4.104)

By substituting the above expressions into Equation 4.99, we obtain


the equations for the amplitudes Ai(z) of the waves. For ­example, to
obtain the equation for the pump beam at ­frequency ω1, we substi-
tute Equation 4.101 into Equation 4.99 to get

dA1 j 3χ( 3 ) k12  4

dz
=−
2β1 |φ1|2 
4 2
 |φ1| |A1| A1 + 2
k =2
∑ 2
|φ1|| φk|2 |Ak|2 A1


+ 2 φ1∗φ2∗φ3 φ4 A2∗ A3 A4 e − j∆βz  , (4.105)
 

where Δβ = (β3 + β4) − (β1 + β2) is the wave vector mismatch.


Performing the power normalization

Ai
Ai → ,
( )
1/2
2ni ε 0 c |φi |2

where ni = neff(ωi) is the effective index of mode i, we can express


Equation 4.105 as (Agrawal 2013)

j 3 χ ( 3 ) k1 
dA1
dz
=−
4ε 0 c 
(
f11 |A1|2 A1 + 2 f12 |A2|2 + f13 |A3|2 + f14 |A4|2 A1 )
+ 2 f1234 A2∗ A3 A4 e − j∆βz  , (4.106)

where

1 |φ1 ||φk |
2 2

f 1k = , (4.107)
n1nk |φ1 |2 |φk |2
Nonlinear Optics Applications of Microring Resonators 187

φ1∗φ2∗φ3 φ4
f1234 = .
4 1/2 (4.108)
∏ n
k =1
k |φk| 
2

If the four frequencies are spaced not too far apart, we can approxi-
mate the waveguide modes as being nearly identical, ϕi ≈ ϕ and
ni ≈ neff, in which case Equation 4.106 simplifies to

dA1
dz
(
= − jγ (ω1 ) |A1|2 A1 + 2 |A2|2 +|A3 |2 +|A4|2 A1 )
+ 2 A2∗ A3 A4 e − j∆ββz  , (4.109)

where
2
3 χ( 3 ) ω  nc  n2ω (4.110)
γ (ω) = = ,
2
4neff ε 0 c 2 Aeff  neff  cAeff
2
|φ|2
Aeff = . (4.111)
|φ|4

The rightmost expression in Equation 4.110 is obtained using the


relation χ( 3 ) = 4nc2 n2 ε 0 c/3, which is found from Equation 4.12. We
also note that the above expressions for the nonlinear coefficient and
effective mode area are the same as those in Equations 4.50 and 4.51.
In a similar manner, we can also derive the equations for A2, A3,
and A4 to get

dA2
dz
( )
= − jγ (ω 2 ) |A2|2 A2 + 2 |A1|2 +|A3|2 +|A4|2 A2 + 2 A1∗ A3 A4 e − j∆βz  ,

(4.112)
dA3
dz
( )
= − jγ (ω 3 ) |A3|2 A3 + 2 |A1|2 +|A2|2 +|A4|2 A3 + 2 A1 A2 A4∗ e j∆βz  ,

(4.113)

dA4
dz
( )
= − jγ (ω 4 ) |A4|2 A4 + 2 |A1|2 +|A2|2 +|A3|2 A4 + 2 A1 A2 A3∗ e j∆βz  .

(4.114)
188 Optical Microring Resonators

In the above equations, the first term in the square brackets on


the right-hand side is responsible for SPM, while the second term
gives rise to XPM effects. The third term is responsible for the
mutual interactions between the four waves, which leads to the
generation and amplification of the new frequency ω4. We can
also include linear propagation loss in the waveguide by adding
the term α0Ai/2 to the left-hand side of Equations 4.109 and 4.112
through 4.114.
Solution of the nonlinear coupled wave equations (4.109) and
(4.112) through (4.114) in general requires a numerical approach.
For the simplified case where linear loss can be neglected and
there is no depletion of the pump beams (|A1|2, |A2|2 ≫ |A3|2,
|A4|2), ­analytical solutions can be obtained. Neglecting XPM effects
induced by the signal and idler waves, the equations for the pump
beams reduce to

dA1
dz
(
= − jγ |A1|2 + 2|A2|2 A1 , ) (4.115)

dA2
dz
(
= − jγ |A2|2 + 2|A1|2 A2 , ) (4.116)

where we have also made the simplification that γ(ω1) ≈ γ(ω2).


The above equations can be solved to give

A1 ( z) = P1 e − jγ ( P1 + 2 P2 )z , (4.117)

A2 ( z) = P2 e − jγ ( P2 + 2 P1 )z , (4.118)

where P1 = |A1(0)|2 and P2 = |A2(0)|2 are the initial powers of the


pump waves. Substituting the above solutions into the equations
for the signal and idler waves and neglecting the SPM terms,
we obtain

dA3
= − j 2γ  2Pavg A3 + P0 A4∗ e ,
j( ∆β − 6 γPavg ) z
(4.119)
dz 

dA4∗
= j 2γ  2Pavg A4∗ + P0 A3 e ,
− j( ∆β − 6 γPavg ) z
(4.120)
dz 
Nonlinear Optics Applications of Microring Resonators 189

where Pavg = (P1 + P2)/2 and P0 = P1P2 are the arithmetic and
­geometric averages, respectively, of the pump powers. Solutions of
Equations 4.119 and 4.120 are given by

A3 ( z) = ( a3 e gz + b3 e − gz )e − jΩz , (4.121)

A4 ( z) = ( a4 e gz + b4 e − gz )e − jΩz , (4.122)

where

g = (2γP0 )2 − ( γPavg + ∆β/2)2 , (4.123)

Ω = 3 γPavg − ∆β/2. (4.124)

The coefficients a3, b3, a4, and b4 in Equations 4.121 and 4.122 can
be determined from the boundary conditions of the signal and
idler waves. The parameter g is called the parametric gain of the
FWM process. From Equation 4.123 we see that gain is achieved
for a range of values of the wave vector mismatch Δβ such that
the term under the square root sign is positive, that is, when
(2γP0)2 > (γPavg + Δβ/2)2.
Of practical interest is the case of degenerate FWM where the
two pump beams A1 and A2 are identical. Relabeling the pump,
signal, and idler waves as Ap, As, Ai, and their frequencies as
­
ωp, ωs, ωi, respectively, we obtain the equations for the three waves
as follows:

dAp
dz
( )
= − jγ (ω p ) |Ap|2 Ap + 2 |As|2 +|Ai|2 Ap + 2 Ap∗ As Ai e − j∆βz  ,

(4.125)

dAs
dz
( )
= − jγ (ω s ) |As|2 As + 2 |Ap|2 +|Ai|2 As + Ap2 Ai∗ e j∆βz  , (4.126)


dAi
dz
( )
= − jγ (ω i ) |Ai|2 Ai + 2 |Ap|2 +|As|2 Ai + Ap2 As∗e j∆βz  , (4.127)

190 Optical Microring Resonators

where Δβ = βs + βi − 2βp. The frequency of the idler wave is given


by ωi = 2ωp − ωs. Under the assumption of no pump depletion,
Equations 4.125 through 4.127 simplify to
dAp
= − jγ|Ap|2 Ap , (4.128)
dz

dAs
dz
(
= − jγ 2|Ap|2 As + Ap2 Ai∗e j∆βz , ) (4.129)

dAi
dz
(
= − jγ 2|Ap|2 Ai + Ap2 As∗e j∆βz , ) (4.130)

where we have also made the approximation γ(ωp) ≈ γ(ωs) ≈ γ(ωi) = γ.


The solution of the pump wave is
− jγPp z
Ap ( z) = Pp e , (4.131)

where Pp = |Ap(0)|2 is the input pump power. The solutions for the
signal and idler waves can be expressed as

As ( z) = [as cosh( gz) + bs sinh( gz)]e − jΩz , (4.132)

Ai ( z) = [ai cosh( gz) + bi sinh( gz)]e − jΩz , (4.133)

where

g = ( γPp )2 − (K/2)2 , (4.134)

Ω = 2γPp − K/2, (4.135)

K = ∆β + 2γPp . (4.136)

From Equations 4.134 and 4.136, we find that gain is achieved for
wave vector mismatch values in the range −4γPp < Δβ < 0. The maxi-
mum achievable gain is gmax = γPp, which occurs when K = 0 or
Δβ = −2γPp.
If we assume that only the signal wave and pump wave are
present at the input of the waveguide, then we have the boundary
conditions |Ai(0)|2 = 0 and |As(0)|2 = Ps, where Ps is the initial power
of the signal wave. In addition, we also have
Nonlinear Optics Applications of Microring Resonators 191

dAs
= − j 2Pp As (0), (4.137)
dz z=0

which can be obtained by evaluating Equation 4.129 at z = 0.


Applying the above boundary conditions to the solutions in
Equations 4.132 and 4.133 and making use of the power conserva-
tion requirement |As(z)|2 + |Ai(z)|2 = Ps, we obtain

As ( z) = Ps [cosh( gz) − ( jK/2 g )sinh( gz)]e − jΩz , (4.138)

 γPp 
Ai ( z) = j Ps   sinh( gz)e − jΩz . (4.139)
 g 

The above solutions describe a power conversion process in


which the idler wave grows exponentially with the propagat-
ing ­distance while the signal wave is being depleted. The power
conversion efficiency after the waves travel a length L of the
­
­waveguide is defined as
2
|Ai (L)|2  γPp 
η= = sinh 2 ( gL). (4.140)
Ps  g 

Another useful parameter characterizing the FWM process is the


coherence length, which is defined as Lc = 2π/|Δβ|. The coherence
length gives the maximum waveguide length over which power
conversion can occur in the presence of wave vector m­ ismatch Δβ
(Agrawal 2013).

4.2 Optical Bistability and Instability


in a Nonlinear Microring Resonator
When an input monochromatic light wave is tuned to a resonant
frequency of a nonlinear microring resonator, the light intensity
in the microring is greatly amplified due to resonance, which can
lead to enhanced nonlinear effects. In a microring resonator with
intensity-dependent refractive index, the simplest manifestation
of nonlinearity is SPM, which causes an intensity-dependent shift
in the resonant frequencies of the microring. At high powers, mul-
tiple solutions of the field inside the resonator can exist for a single
192 Optical Microring Resonators

input power value. Some of these solutions may be unstable, which


give rise to nonlinear processes such as optical bistability, self-­
oscillations, and even chaos. In addition to the important funda-
mental physics that they entail, these n ­ onlinear processes are also
of practical interest for implementing optical switches, threshold-
ers, flip-flops, and memory elements.
In this section we study the behavior of a microring resona-
tor with nonlinear refractive index under continuous-wave (CW)
excitation. We begin in Section 4.2.1 with the analysis of SPM in an
APMR with an intensity-dependent refractive index. Section 4.2.2
shows how optical bistability and instability in the microring can
arise due to instantaneous Kerr nonlinearity at moderate to high
powers. The influence of free carriers and the effects due to the
finite response time of FC-induced nonlinearity will be discussed
in Section 4.2.3.
In general, the analysis of the nonlinear dynamics in a microring
resonator can be carried out in either the framework of energy cou-
pling or power coupling. However, since the energy coupling formal-
ism is valid only over a limited frequency range around a r­ esonance,
it is inadequate for studying the nonlinear behavior of the resonator
at large frequency detunings and high powers. An important exam-
ple of the limitations of the energy coupling ­formalism is that it fails
to predict Ikeda instability (Ikeda et al. 1980), which is caused by
the generation and mixing of adjacent cavity modes (Silberberg and
Bar-Joseph 1984). In Sections 4.2.1 and 4.2.2 below, we will adopt the
power coupling formalism to ­analyze SPM in ­microring ­resonators
with instantaneous n ­ onlinearity. On the other hand, since the energy
coupling formalism provides a more straightforward approach for
incorporating nonlinearity with finite response time, we will use it
to obtain an approximate analysis of the resonator dynamics in the
presence of free carriers in Section 4.2.3.

4.2.1 SPM and enhanced nonlinearity


in a microring resonator
We consider an APMR with radius R and field coupling coefficient
κ as shown in Figure 4.2a. The microring waveguide is assumed
to have effective index nr, linear loss coefficient α0, and instanta-
neous intensity-dependent refractive index in the waveguide core
with Kerr coefficient n2. The input port of the APMR is excited
Nonlinear Optics Applications of Microring Resonators 193
(a) (b)
Low Pin
High Pin
R RHS

A(z)
z=0

sin κ, τ sout LHS

0 –φNL

Figure 4.2 (a) Schematic of an APMR for nonlinear dynamic analysis


and (b) graphical solutions of Equation 4.145 showing a single real solu-
tion for low input power and multiple real solutions for high input power.
The LHS curve is the left-hand side expression of Equation 4.145 and the
straight lines are the right-hand side (RHS) expression.

by a CW signal sin with frequency ω and power Pin = |sin|2. We


denote the field circulating inside the microring by A(z), where
z is the coordinate along the microring circumference with the
origin (z = 0) located just after the coupling junction, as indicated
in Figure 4.2a. Near a microring resonance, the frequency dis-
persion due to resonance is much larger than the dispersion of
the microring waveguide, so we can neglect the latter and write
the equation for the nonlinear propagation of the field inside the
microring as

dA α 0
+ A = − jγ |A|2 A, (4.141)
dz 2

where γ is the nonlinear coefficient defined in Equation 4.51. The


solution of the above equation is given by Equation 4.60,

A( z) = A(0)e − α0 z/2 e − jγP0 zeff , (4.142)

where P0 = |A(0)|2 is the power in the microring at z = 0, and


zeff = (1 − e − α0 z )/α 0 is the effective propagation distance. At the
coupling junction between the microring and the waveguide, the
­following boundary condition must be satisfied:

A(0) = τA(L)e jφL − jκ Pin , (4.143)


194 Optical Microring Resonators

where τ = 1 − κ 2 is the transmission coefficient of the coupling


junction and ϕL = −nr(ω/c)2πR is the linear round-trip phase of the
microring. Applying the above boundary condition to the solution
in Equation 4.142, we solve for A(0) to get

− jκ Pin
A(0) = . (4.144)
1 − τart e j( φL + φNL )

In the above expression, art = e −α0 πR is the linear round-trip ampli-


tude attenuation, ϕNL = −γP0Leff is the nonlinear round-trip phase,
and Leff = (1 − e −2α0 πR )/α 0 is the effective length of the microring cir-
cumference. Taking the absolute square of Equation 4.144, we obtain
an equation for ϕNL which can be expressed as
1 φNL
=− , (4.145)
1 + F sin 2 ((φL + φNL )/2) γPr,max Leff

where F = 4τart/(1 − τart)2 is the contrast factor defined in Equation


2.46, and Pr,max is the maximum power in the microring at resonance,

κ 2 Pin
Pr,max = . (4.146)
(1 − τart )2

Equation 4.145 can be solved using an iterative numerical technique


to obtain the nonlinear phase shift ϕNL and the power in the micror-
ing (P0 = −ϕNL/γLeff) for a given input power Pin and linear phase
detune ϕL. Graphically, we can depict the solutions for ϕNL as the
intersections between the resonance curve given by the left-hand
side of Equation 4.145 and the straight line given by the right-hand
side, as shown in Figure 4.2b. We find that, for low input powers,
Equation 4.145 has only one real solution. At high powers, however,
multiple real solutions can occur for a given input power and linear
phase detune. Some of these solutions turn out to be unstable and
we will discuss them in more detail in Section 4.2.2. For low input
powers such that the nonlinear phase shift is small, we can obtain
an approximate solution to Equation 4.145 by linearizing the sine
function around ϕL. This results in the cubic polynomial

{1 + F[sin(φ /2) + (φ
L NL }
/2)cos(φL /2)]2 φNL = − γPr,max Leff . (4.147)

The solutions for ϕNL are taken as the real roots of the above
polynomial.
Nonlinear Optics Applications of Microring Resonators 195

Neglecting the small nonlinear phase shift ϕNL at low input


powers, Equation 4.144 indicates that the power in the microring
resonator is enhanced by approximately the square of the linear
field enhancement factor,
2
P0 − jκ
≈ = FE2 . (4.148)
Pin 1 − τart e jφL

As a result, we expect nonlinear effects in the microring to be also


enhanced by a factor FE2 due to resonance. On the other hand, SPM
causes the microring resonant frequency to shift by an amount
ΔωNL given by (see Equation 4.62)

∆ω NL n φ n γP
= −  r  NL ≈ − r 0 , (4.149)
ω n
 g Lφ ng βr

where βr = nr(ω/c) is the linear propagation constant in the micro­


ring. Equation 4.149 indicates that the amount of self-detuning is
­proportional to the power in the microring. Thus, in order to achieve
maximum enhancement of nonlinear effects in the r­esonator,
­
we must initially detune the frequency ω of the input signal from
the linear resonance by ΔωNL.
Figure 4.3 illustrates the effects of SPM in a silicon APMR. The
microring has radius R = 20 μm and coupling coefficient κ = 0.1.
(a) (b)
25
Power enhancement, P0/Pin (dB)

Nonlinear phase shift,φNL/π

20 20 0.02
Normalized power (dB)

P0/Pin
15
10 15 0.015
P0/Pin
5
0 10 0.01

–5
Pout/Pin 5 0.005
–10 φNL
–15
0 0
–0.04 –0.03 –0.02–0.01 0 0.01 0.02 0.03 0.04 10–5 10–4 10–3 10–2 10–1 100
Linear phase detune, φL/π Input power, Pin (W)

Figure 4.3 Self-phase modulation effects in a nonlinear Si APMR: (a)


spectral responses of the output power (Pout) and the power in the micro­
ring (P0) normalized to a fixed input power Pin = 5 mW. Solid lines are
responses in the presence of Kerr nonlinearity; dashed lines are responses
of the linear device. (b) Power enhancement in the microring (P0/Pin) and
self-phase detuning (|ϕNL|/π) as functions of the input power when the
applied CW signal is initially tuned to a microring resonance (ϕL = 0).
196 Optical Microring Resonators

The microring waveguide is assumed to have effective index nr = 2.5,


linear loss α0 = 2.5 dB/cm, Kerr coefficient n2 = 4.5 × 10−14 cm2/W,
and effective mode area Aeff = 0.15 μm2 near the 1.55 μm wave-
length. FCD and FCA are neglected. Figure 4.3a plots the power
in the microring (P0) and the power at the output port (Pout) of the
APMR as functions of the linear phase detune ϕL for a fixed input
power Pin = 5 mW. Also shown for comparison are the spectral
responses of the power in the microring and the output power of
a linear APMR. The shift in the resonance of the microring due to
SPM is apparent from the plot. In addition, we observe that SPM
causes the shapes of the spectral responses of the nonlinear APMR
to skew to the right (i.e., toward lower frequencies), resulting in a
steepening of the right edge of the spectra. Further increase in the
input power will lead to bistable behavior, which will be discussed
in the next section. Figure 4.3b shows the power enhancement in
the microring (P0/Pin) and the nonlinear phase shift (|ϕNL|) as func-
tions of the input power when the applied CW signal is initially
tuned to a microring resonance (ϕL = 0). We observe that as the
input power is increased, the r­ esonant frequency detuning due to
SPM also increases, which in turn leads to a reduction in the power
enhancement in the microring.

4.2.2 Bistability and self-pulsation in a microring resonator


We saw from the plot in Figure 4.2b that at moderate to high
input powers, Equation 4.145 admits multiple real solutions of the
­nonlinear round-trip phase for a given input power and ­linear
phase detune. Some of these solutions may not be stable and,
depending on the nature of the instability, can lead to bistable
behavior or ­oscillation of the field inside the microring. To a­ nalyze
the nonlinear dynamics in the APMR near an unstable solution, we
rewrite the b
­ oundary condition (4.143) at the c­ oupling ­junction of
the microring in a slightly different way:
2
A(0, t + Trt ) = τart A(0, t)e jφL e − jξ|A( 0 ,t )| − jκ Pin , (4.150)

where A(0, t) is the field in the microring at z = 0 and time t, and


A(0, t + Trt) is the field at the same point but delayed by one round-
trip time Trt. For convenience, we have also defined a new ­nonlinear
parameter, ξ = γLeff. Equation 4.150 describes the time evolution
Nonlinear Optics Applications of Microring Resonators 197

of the field A in the microring in discrete intervals of Trt. In the


­theory of nonlinear dynamics, such an equation is called an ­iterated
map. In particular, Equation 4.150 is a 2D map since A is a complex
quantity. For a given set of parameters, the iterated map generates
a sequence of field values An = A(0, t + nTrt), which may converge
to a fixed value, oscillate about it, or completely diverge from it.
The fixed points (or stationary points) of the iterated map are deter-
mined by setting A(0, t + Trt) = A(0, t) = A0 in Equation 4.150 and
­solving for A0. The result is

− jκ Pin
A0 = 2 , (4.151)
1 − τart e jφL e − jξ|A0|

which is the same as Equation 4.144. The fixed-point power in the


microring, P0 = |A0|2, can thus be determined by solving Equation
4.145. The phase of A0 is then evaluated by substituting |A0|2 = P0
into Equation 4.151.
To determine the dynamic behavior of the field around a fixed-
point solution A0, we perform linear stability analysis by adding a
small perturbation ε(t) to the solution,

A(t) = A0 + ε(t). (4.152)

Substituting the above expression for A(t) into Equation 4.150 and
making the approximation

{ }
2 2
e − jξ|A0 + ε(t )| ≈ e − jξ|A0| 1 − jξ[ A0∗ ε(t) + A0 ε∗ (t)] ,

which is valid for small perturbation ε(t), we get

{ } 2
A0 + ε(t + Trt ) = τart e jφL [ A0 + ε(t)] 1 − jξ  A0∗ ε(t) + A0 ε∗ (t) e − jξ|A0|
− jκ Pin . (4.153)

Keeping only linear terms in ε(t) and making use of Equation 4.151,
we can further simplify the above expression to

( )
ε(t + Trt ) = G  1 − jξ|A0|2 ε(t) − jξA02 ε∗ (t) , (4.154)
198 Optical Microring Resonators
2
where G = τart e jφL e − jξ|A0| is the round-trip phase and attenuation fac-
tor. Equation 4.154 and its complex conjugate can be put in matrix
form as

 ε(t + Trt )  G(1 − jξ|A0|2 ) − jξGA02   ε(t)   ε(t) 


 ε∗ (t + T ) =  2  ∗  ≡ M  ∗ .
 rt 
∗ ∗ 2
 jξG ( A0 ) G (1 + jξ|A0| )  ε (t)

 ε (t)
(4.155)

The above equation describes the discrete time evolution of the per-
turbation ε(t). Writing the solution as

 ε(t + Trt )  σTrt  ε(t) 


 ε∗ (t + T ) = e  ε∗ (t) , (4.156)
 rt   

and substituting it into Equation 4.155, we get

G(1 − jξ|A0|2 ) − λ − jξGA02   ε(t) 


   ∗  = 0, (4.157)
 jξG∗ ( A0∗ )2 G (1 + jξ|A0| ) − λ   ε (t)
∗ 2

where λ = e σTrt is identified as an eigenvalue of the matrix M. The


solution for the perturbation ε after n round-trips is given by

ε(t + nTrt ) = λ n ε(t). (4.158)

It is apparent from the above result that the behavior of the per-
turbation depends on the eigenvalues λ. If both eigenvalues have
magnitude less than 1, the perturbation will eventually die off and
the perturbed solution A(t) converges to the fixed-point value A0. In
this case we obtain a stable field inside the microring. On the other
hand, if either eigenvalue of M has magnitude greater than 1, the
solution is unstable. The nature of the instability depends on the
phase φ of the eigenvalue, λ = |λ|ejφ. Specifically, if the eigenvalue
is real and greater than 1 (φ = 0 and |λ|> 1), the perturbed solution
A(t) will be repelled from the fixed-point value A0, to be attracted
to nearby fixed points. This situation corresponds to optical bista-
bility. If the eigenvalue is real and less than −1 (φ = π and |λ| > 1),
the perturbation changes sign every round-trip time Trt according
to Equation 4.158. The field in this case oscillates around the fixed
point with a period exactly equal to twice the microring round-trip
time. This is called period-doubling oscillation or Ikeda instability
Nonlinear Optics Applications of Microring Resonators 199

(Ikeda et al. 1980). For other values of φ (λ is complex with |λ| > 1),
the solution oscillates with a period equal to 2πTrt/φ. This behavior
is referred to as self-pulsation.
For the matrix M in Equation 4.155, the eigenvalues can be
­evaluated to give

( )
λ ± = τart θ ± θ 2 − 1 , (4.159)

θ = cos(φL − ξP0 ) + ξP0 sin(φL − ξP0 ). (4.160)

The parameter θ depends on the linear phase detune ϕL and the


power P0 in the microring. If θ2 < 1, the eigenvalues are complex with
magnitude |λ±| = τart < 1, so the solution is stable. If θ2 ≥ 1, the eigen-
values are real. In this case bistability occurs if θ > (τart + 1/τart)/2
whereas Ikeda instability occurs if θ < −(τart + 1/τart)/2. Since the
eigenvalues are never complex with magnitude greater than 1, self-
oscillation other than the Ikeda type cannot occur in a microring
with instantaneous Kerr nonlinearity. We will see in Section 4.2.3,
however, that if the nonlinearity has a finite response time, such as
due to Debye relaxation or free carrier diffusion and recombination,
self-pulsation can occur with oscillation periods in the order of the
relaxation time (Ikeda et al. 1982, Silberberg and Bar-Joseph 1984).
Ikeda instability typically occurs at very high input ­powers and
large linear phase detunes. It is characterized by self-­oscillations at
a period exactly equal to twice the microring round-trip time (2Trt),
and this period is insensitive to changes in the phase detune or
input power as long as the condition θ < −(τart + 1/τart)/2 is met. Thus,
there exist certain ranges of ϕL and Pin values for which period-dou-
bling oscillations can occur. The origin of period-­doubling oscilla-
tions can be explained in terms of the spontaneous generation of
the two cavity modes adjacent to the input frequency due to FWM
(Silberberg and Bar-Joseph 1984). Suppose an input CW wave tuned
to a frequency ω0 about half way between two resonance modes
ωm and ωm+1 of the microring, as shown in Figure 4.4. As the input
power is increased, the cavity modes are shifted to lower frequen-
cies due to SPM (assuming n2 > 0). When the input frequency falls
exactly half way between the two modes, spontaneous FWM gener-
ates two new frequencies at the cavity modes, which then beat with
the input signal to give rise to ­oscillations at a period equal to 2/
FSR, or twice the microring round-trip time.
200 Optical Microring Resonators

ωm ωm+1
Low
power

ω
High
FSR/2
power

ωm ω0 ωm+1 ω

Figure 4.4 Diagram depicting the origin of period-doubling oscilla-


tions due to spontaneous FWM generation of the cavity modes ωm and
ωm+1 adjacent to the input frequency ω0. (After Silberberg, Y., Bar-Joseph, I.,
1984, J. Opt. Soc. Am. B, 1, (4): 662–670.)

As a numerical example illustrating SPM effects in a micror-


ing, we show in Figure 4.5 the stability curves of a silicon APMR
for two sets of coupling coefficient and linear phase detune values:
(a) κ = 0.1 and ϕL = 0.025π, and (b) κ = 0.7 and ϕL = 0.5π. The radius
of the microring is R = 100 μm. The microring waveguide has
effective index nr = 2.5, linear loss α0 = 2.5 dB/cm, Kerr coefficient
n2 = 4.5 × 10−14 cm2/W and effective mode area Aeff = 0.15 μm2 near
the 1.55 μm wavelength. Both FCD and FCA are neglected. The
plots show the fixed-point solutions of the power in the microring
(P0) as a function of the input power (Pin). The regions of stability
(S), bistability (BS) and Ikeda oscillations (IK) are indicated on the
plots. We see that bistability can occur at moderate input powers
and small linear phase detunes. On the other hand, the threshold
power required to reach Ikeda instability is significantly higher and
is typically beyond the power levels that can be handled by most
integrated optics waveguides. For this reason, Ikeda instability has
not been observed in a microring resonator.

4.2.3 Free carrier effects and self-pulsation


in a microring resonator
In the previous section, we saw that in a microring resonator with
instantaneous Kerr nonlinearity, period-doubling oscillations can
occur at very high powers and large linear phase detunes. In this
section, we show that if the nonlinearity has a finite response time,
self-pulsation can occur at much lower powers. An example of where
this behavior may be observed is in a semiconductor microring
resonator with a significant concentration of free carriers generated
Nonlinear Optics Applications of Microring Resonators 201

(a) 0.8

1
Power in microring, P0 (W) 0.6
0.8

Pin/ Pout
0.6

0.4 0.4
BS 0.2
0 0.02 0.04 0.06 0.08 0.1
Pin (W)
0.2

0
0 0.02 0.04 0.06 0.08 0.1
Input power, Pin(W)

(b) 20
IK

S
15
Power in microring, P0 (W)

10

BS

0
0 5 10 15 20 25
Input power, Pin(W)

Figure 4.5 Plots of the power in a Si APMR (P0) versus input power: (a)
κ = 0.1 and ϕL = 0.025π, (b) κ = 0.7 and ϕL = 0.5π. Regions of stability, bista-
bility, and Ikeda instability are indicated by S, BS, and IK, respectively.
Inset of (a) shows a plot of the power transmission of the APMR (Pout/Pin)
versus the input power.
202 Optical Microring Resonators

from TPA. We will find that in such a resonator, the period of oscil-
lation depends on the FC lifetime and that there is an upper limit to
this value for which self-pulsation can occur (Armaroli et al. 2011,
Malaguti et al. 2011).
Analysis of the nonlinear dynamics of a microring resonator
with FC-induced nonlinearity is complicated by the addition of a
differential equation governing the evolution of the carrier density.
Since we are mainly interested in the behavior of the resonator in the
vicinity of a resonance (i.e., small linear phase detunes), it is more
expedient to employ the energy coupling formalism for our analysis
(Malaguti et al. 2011, Chen et al. 2012). In this model, the nonlinear
dynamics of the resonator are described by a system of continuous-
time differential equations rather than the discrete-time iterated
map in Equation 4.150. It should be kept in mind, however, that the
energy coupling formalism only provides an approximate model of
the resonator. We will discuss the limitations of this model in pre-
dicting the nonlinear dynamic behavior of a microring resonator.
We consider again the APMR in Figure 4.2a, which is excited by
an input monochromatic wave at frequency ω with power Pin. Let
a(t) represent the amplitude of the wave in the microring, which is
­normalized so that |a(t)|2 gives the total energy in the resonator.
The rate of energy coupling between the straight waveguide and
the microring is denoted by μ, which is related to the field coupling
­coefficient κ by μ = κ/Trt, where Trt = 2πR/vg is the microring round-
trip time and vg is the group velocity. We assume that the micror-
ing waveguide possesses instantaneous Kerr nonlinearity as well
as FC-induced nonlinearity. Including both the effects of ­nonlinear
dispersion and nonlinear absorption, the equation ­governing the
­evolution of the energy amplitude in the microring can be expressed as

da
= − j(∆ω − ∆ω NL )a − ( γ L + γ NL )a − jµ Pin , (4.161)
dt

where Δω = ω − ω0 is the detuning from the resonant frequency


ω0 of the microring, ΔωNL is the nonlinear frequency shift, and
γL and γNL are the linear and nonlinear decay rates, respectively.
The ­nonlinear frequency shift of the microring resonance is related
­ onlinear index change ΔnNL by
to the n

∆nNL
∆ω NL = −ω 0 , (4.162)
ng
Nonlinear Optics Applications of Microring Resonators 203

where ng is the group index of the microring waveguide. We can


determine the nonlinear index change by summing the contribu-
tions from Kerr nonlinearity and FCD,
2
 n  n2 |a(t)|2  nc 
∆nNL = ∆nNL,K + ∆nNL,fc =  c  +   σ r N(t), (4.163)
 nr  Aeff Trt  nr 

where nc is the refractive index of the waveguide core, nr the effec-


tive index, n2 the Kerr coefficient, Aeff the effective mode area, N(t)
the FC density and σr is the FC refraction volume. The FC density in
the microring is given by Equation 4.80,
2 4
dN N  nc  α2 a(t)
(4.164)
+ =  ,
dt τ fc  nr  2ωAeff Trt2
2

where α2 is the TPA absorption coefficient. In evaluating the Kerr


nonlinear index change in Equation 4.163, we have approximated
the power in the microring as |a(t)|2/Trt.
The decay rates γL and γNL in Equation 4.161 are related to the
linear and nonlinear loss, respectively, in the microring. Specifically,
the linear decay rate can be computed from the linear loss coefficient
α0 of the microring waveguide and the coupling coefficient μ as

α 0 vg µ 2
γL = + . (4.165)
2 2

The nonlinear decay rate is due to TPA and FCA and can be evalu-
ated using the nonlinear absorption coefficients in Equations 4.69
and 4.70,

α NL ,K vg α NL ,fc vg vg  nc  α 2 |a(t)|2  nc  
2

γ NL = + =   +   σ a N(t) ,
2 2 2  nr  Aeff Trt  nr  
(4.166)
where σa is the FCA cross section. Substituting the above results
for the nonlinear frequency shift ΔωNL and nonlinear decay rate γNL
into Equation 4.161, we can summarize the equations for the energy
amplitude and FC density in the microring as follows:

da
= −( j∆ω + γ L )a − jηK |a|2 a − jηfc Na − jµ Pin , (4.167)
dt
204 Optical Microring Resonators

dN N
+ a 4,
= ς|| (4.168)
dt τ fc

where
2
n   jα 2  vg
ηK = ηK′ − jηK′′ =  c   n2 k0 − ,
 nr   2  Aeff Trt

nc  jσ a 
ηfc = ηfc′ − jηfc′′ =  σ r k0 − 2  vg ,
nr
2
n  α2
ς= c .
 nr  2ωAeff
2
Trt2

Along with Equation 4.168, Equation 4.167 and its complex con-
jugate form a 3D nonlinear dynamical system. The fixed points a0
and N0 of the system are determined by setting da/dt = 0 and dN/
dt = 0 to get
N 0 = τ fc ς|a0|4 , (4.169)

− jµ Pin
a0 = . (4.170)
j∆ω + jηK |a0|2 + jηfc N 0 + γ L

To determine the stability of the fixed points, we add a small per-


turbation to a0 and N0 to get a(t) = a0 + ε(t), N(t) = N0 + δn(t). Making
these substitutions in Equations 4.167 and 4.168 and keeping only
first-order terms in ε(t) and δn(t), we obtain

= −( j∆ω + j 2ηK |a0 |2 + jηfc N 0 + γ L )ε − jηK |a0 |2 ε∗ − jηfc a0 δn,
dt
(4.171)
dδn δn
= 2ς|a0 |4 ( a0∗ ε + a0 ε∗ ) − . (4.172)
dt τ fc

The above equations can be put in the matrix form, dε/dt = Mε,
where ε = [ε(t), ε*(t), δn(t)]T and the matrix M is given by
 −Γ − jηK a02 − jηfc a0 
 
M =  jη∗K ( a0∗ )2 −Γ ∗ jη∗fc a0∗  , (4.173)
 2ς|a0|2 a0∗ 2ς|a0|2 a0 −1/τ fc 

Nonlinear Optics Applications of Microring Resonators 205

Γ = j∆ω + j 2ηK |a0|2 + jηfc N 0 + γ L .

It is apparent that the time evolution of the perturbation vector ε is


determined by the eigenvalues of the matrix M. In particular, if all
the eigenvalues reside in the left half of the complex plane, the fixed-
point solutions a0 and N0 are stable. If one of the eigenvalues crosses
into the right half of the complex plane, the fixed point becomes
unstable. In this case bistability occurs if the e­igenvalue is real
and positive, and Hopf bifurcation occurs if the eigenvalues become
complex with positive real parts. The latter case ­corresponds to self-
pulsation with the period of oscillation equal to |2π/Im{λ}|.
Before considering the behavior of the APMR in the presence
of free carriers, we first look at the simple case where only instan-
taneous Kerr nonlinearity is present (ς = 0, ηK′′ = 0, ηfc = 0). From
Equation 4.170 we obtain the following cubic polynomial for the
energy E0 = |a0|2 in the microring:

(δ + hE0 )2 + 1 E0 = Ein,c , (4.174)

where δ = Δω/γL is the normalized frequency detune, h = ηK′ /γ L the


nonlinear factor and Ein,c = μ2Pin/γL is the energy coupled into the
microring. The above polynomial has a single real root for δ < 3.
For δ > 3, three real roots exist for microring energies in the range

2δ + δ 2 − 3 2δ − δ 2 − 3 (4.175)
− < E0 < − .
3h 3h

To evaluate the stability of these roots, we determine the eigenval-


ues of the matrix M in Equation 4.173, which in this case reduces to

 −Γ − jηK′ a02 
M= ∗ 2 , (4.176)
 jηK′ ( a0 ) −Γ ∗ 

where Γ = j(∆ω + 2ηK′ E0 ) + γ L. We find that the normalized


­eigenvalues, λn = λ/γL, of the above matrix are given by

λ n = −1 ± j (δ + hE0 )(δ + 3 hE0 ). (4.177)

For microring energies in the range given by Equation 4.175, the


eigenvalues are real with one being negative and the other positive.
Thus the two limits in Equation 4.175 correspond to the lower and
206 Optical Microring Resonators

upper energy thresholds for observing bistable behavior. Outside


this range of energies, the eigenvalues can be real or complex but
the real parts are always negative, so the solutions are always sta-
ble. Note that the above analysis fails to predict instability of the
Ikeda type. This is due to the fact that the energy coupling model
neglects neighboring resonance modes of the microring, whose
interactions give rise to period-doubling oscillations as discussed
in Section 4.2.2.
We next consider the effects of free carriers on the nonlinear
dynamics of the APMR (Armaroli et al. 2011, Malaguti et al. 2011).
For simplicity, we will assume that FCD dominates over Kerr non-
linearity so that we can set ηK = 0. In addition, we will also neglect
FCA ( ηfc′′ = 0) in the analysis although it can have a significant
impact on the threshold powers for observing instability, as will
be discussed in the example at the end of the section. With these
simplifications, we find that the fixed-point values of the energy in
the microring are given by the roots of the fifth-order polynomial

(δ + hE02 )2 + 1 E0 = Ein,c , (4.178)

where δ = Δω/γL, Ein,c = μ2Pin/γL and the nonlinear factor is given


by h = ηfc′ τ fc ς/γ L. The above polynomial has multiple real roots for
microring energies in the range E0−, bs < E0 < E0+, bs, where
1/2
±
 3δ ± 4δ 2 − 5 
E0 , bs = −  . (4.179)
 5h 

The stability of these roots are determined by evaluating the eigen-


values of the matrix M, which in this case simplifies to

 −Γ 0 − jηfc′ a0 
 
M= 0 −Γ ∗ jηfc′ a0∗  , (4.180)
 2 ∗ 2 
 2ς a0 a0 2ς a0 a0 −1/τ fc 

where Γ = j(∆ω + ηfc′ τ fc ςE02 ) + γ L. The normalized eigenvalues,


λn = λ/γL, of the above matrix are found to be solutions of a cubic
­polynomial of the form

λ 3n + a2 λ n2 + a1λ n + a0 = 0, (4.181)
Nonlinear Optics Applications of Microring Resonators 207

with the coefficients given by

a2 = 2 + 1/τ n ,
2
a1 = 1 + 2/τ n + δ NL ,
2
a0 = (1 + δ NL + 4 hE02δ NL )/τ n .

In the above expressions, τn = τfcγL is the normalized FC lifetime


and δ NL = δ + hE02 is the normalized nonlinear frequency detune.
The normalized FC lifetime can also be interpreted as the ratio of the
FC lifetime to the linear cavity lifetime, τ n = τ fc /τ L′ , where τ L′ = 1/γ L.*
For microring energies in the range E0−, bs < E0 < E0+, bs, where E0,bs ±
are
given by Equation 4.179, it can be shown that one of the eigenval-
ues obtained from Equation 4.181 is real and positive, indicating
that bistability will occur in this range. Outside this energy range,
the eigenvalues can become complex with ­positive real part, which
signifies self-pulsation behavior. The threshold energies for the
onset of self-pulsation are determined by the Hopf bifurcation
points, which are defined as the points where a complex eigenvalue
crosses the imaginary axis into the right half-plane. These points
can be found using the Routh–Hurwitz stability ­criterion for a cubic
­characteristic polynomial, a1a2 = a0. We find that, in general, there
are two Hopf bifurcation points occurring at the energies:
1/2
±
 −(1 − τ n )δ ± ∆ 
E 0 , sp =  , (4.182)
 ( 2 − τ n )h 

∆ = (δ 2 + 3) − (τ 2n − 2/τ n ). (4.183)

Self-pulsation occurs for microring energies in the range


E0−,sp < E0 < E0+,sp. If E0,sp
±
are complex, self-pulsation cannot occur.
We can find the condition for which self-pulsation ceases to
occur by setting Δ = 0, yielding the equation

τ 3n,c − (δ 2 + 3)τ n,c − 2 = 0. (4.184)

For a given frequency detune δ, solution of the above equation gives


the critical value τn,c, which defines the upper limit of the normalized

* Here we define the cavity lifetime τ ′L as the time it takes the wave amplitude a(t),
not the energy |a(t)|2, in the microring to decay to 1/e of its initial value.
208 Optical Microring Resonators
5.5

Maximum FC lifetime, τn,c


4.5

4 No SP
τn > τn,c
3.5

3 SP
τn < τn,c
2.5

2
0 1 2 3 4 5
Normalized frequency detune, δ

Figure 4.6 Plot of the maximum normalized FC lifetime for which


self-pulsation can occur versus the normalized frequency detune.
Self-pulsation occurs in the region where τn < τn,c.

FC lifetime for which self-pulsation can occur. A plot of the criti-


cal value τn,c versus the normalized frequency detune is shown in
Figure 4.6. Self-pulsation can occur only in the region τn < τn,c. From
the plot, we find that in an APMR where FC-induced nonlinearity
dominates, self-pulsation is always possible if τn < 2.
As a numerical example, we consider a silicon APMR around
the 1.55 μm wavelength where FCD is assumed to be the domi-
nant source of nonlinearity. The microring has radius R = 25 μm,
effective index nr = 2.5, linear loss α0 = 2.5 dB/cm, and field cou-
pling coefficient κ = 0.1. Assuming a group index ng = 4.5 for the
waveguide, we calculate the corresponding energy coupling coef-
ficient to be μ = 6.5 × 104 s−1/2 and linear decay rate γL = 4 GHz (or
cavity lifetime τ L′ = 0.25 ns). The silicon waveguide has TPA coef-
ficient α2 = 0.8 cm/GW, FC refraction volume σr = −5.3 × 10−21 cm3,
and effective mode area Aeff = 0.15 μm2. We neglect Kerr-induced
dispersion and FCA effects. In Figure 4.7a we plot the power in the
microring (P0 = E0/Trt) and the output power of the APMR (Pout) as
functions of the linear frequency detune for input power Pin = 1 mW
and normalized FC lifetime τn = 2. Also shown for comparison are
the spectral responses of the device in the absence of any nonlinear-
ity. Since free carriers induce a negative refractive index change in
the microring waveguide, the spectral responses of the device are
skewed towards positive frequency detunes as the input power is
increased. In Figure 4.7b we plot the power in the microring (P0) ver-
sus the input power (Pin) for frequency detune δ = 2 and normalized
Nonlinear Optics Applications of Microring Resonators 209
(a) 25 (b) 0.06
20 S

Power in microring, P0 (W)


0.05
15 P0/Pin SP
Normalized power (dB)

10 0.04
5
0.03 BS
0
–5
0.02
–10 Pout/Pin
–15 0.01 S δ=2
–20 τn = 2
–25 0
–20 –15 –10 –5 0 5 10 15 20 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Normalized frequency detune, δ Input power, Pin (mW)
(c) 0.25 (d) 0.25
0.1 0.1
δ=2 τn = 2
Power in microring, P0 (W)

0.1
Power in microring, P0 (W)
0.2 0.2 SP
0.15 0.15
SP
0.2 0.2
+
0.15 P0 , sp 0.15
0.3 0.3
0.4 0.4
0.1 – 0.1 0.6 0.6
P0 , sp 0.8
τn,c 0.8 1.2
+
P0 , bs 1.2 2
BS 2
0.05 0.05 0
S 0 BS
S – 0
P0 , bs
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3
Normalized FC lifetime, τn Normalized frequency detune, δ

Figure 4.7 (a) Plots of the output power (Pout) and power in the micror-
ing (P0) versus the linear frequency detune (δ) for Pin = 1 mW and τn = 2.
Solid lines are responses in the presence of TPA and FCD; dashed lines
are responses in the linear case. (b) Plot of the power in the microring (P0)
versus the input power (Pin) for δ = 2.0 and τn = 2, showing regions of sta-
bility (S), bistability (BS), and self-pulsation (SP). (c) Threshold powers for
bistability and self-pulsation as functions of the normalized FC lifetime
τn at a fixed δ = 2.0. (d) Stability map of microring power versus frequency
detune showing regions of stability, bistability, and self-pulsation for a
APMR with τn = 2. Values of the contour lines in the SP region are normal-
ized periods of oscillations, Tsp /τ′L.

FC lifetime τn = 2. The regions of stability (S), bistability (BS), and


self-pulsation (SP) are indicated on the plot. Compared to the results
in Figure 4.5 where FC effects are neglected and i­ nstantaneous Kerr
nonlinearity dominates, the thresholds for bistability in this case
occur at much lower powers since FC-induced nonlinearity is much
stronger than the Kerr effect in silicon. In addition, self-pulsation
can also be observed at relatively low input powers. Note, however,
that we have assumed an FC lifetime τfc of only 2τ L′ , or about 0.5 ns,
which is smaller than typically observed in a silicon waveguide.
In addition, nonlinear loss due to FCA has also been neglected.
210 Optical Microring Resonators

To investigate the influence of the FC lifetime on the nonlin-


ear dynamics of the microring, we plot in Figure 4.7c the thresh-
old powers for bistability (P0±, bs = E0±, bs /Trt ) and self-pulsation
(P0±, sp = E0±, sp /Trt ) as functions of the normalized FC lifetime for a
fixed frequency detune δ = 2. We find that for any value of τn, there
always exists a range of powers in which bistability will occur. On
the other hand, self-pulsation occurs only for values of τn up to
a critical value of 2.8, or a maximum FC lifetime of about 0.7 ns.
Figure 4.7d shows the regions of frequency detunes and micror-
ing powers for which the device exhibits stable, bistable, and self-
pulsation behaviors. The normalized FC lifetime is set to be τn = 2.
In the SP region, c­ ontour lines are also drawn whose values indicate
the normalized period of oscillation, Tsp /τ L′ . We observe that at low
powers, the pulsation period is typically in the order of the cavity
lifetime.
When the effects of Kerr-induced nonlinearity and FCA are
also taken into account, the nonlinear behaviors of the Si APMR
are noticeably altered, as shown in Figure 4.8. In the simulations,
we assume a Kerr coefficient n2 = 4.5 × 10−14 cm2/W and FCA cross
section σa = 1.45 × 10−17 cm2 for the silicon waveguide. Figure 4.8a
shows the stability curve of the APMR for frequency detune δ = 2
and normalized FC lifetime τn = 2. For comparison, the stability
curve obtained when Kerr nonlinearity and FCA are neglected
is also shown in the plot. In general, Kerr-induced index change
has negligible effect on the nonlinear dynamics of the microring
since it is much smaller than FCD. On the other hand, FCA can
significantly increase the cavity loss, so higher powers are required
to reach bistable and self-pulsation regimes. Another effect of FCA
is that it reduces the critical FC lifetime value τn,c and the range
of microring powers for which self-pulsation can occur. This can
be seen by comparing the plots in Figures 4.7c and 4.8b, which
show the threshold powers for bistability and self-pulsation versus
τn for a fixed δ = 2. Figure 4.8c shows the stability map of micror-
ing power versus frequency detune for τn = 2. Again, the region of
self-­pulsation ­behavior is much reduced compared to the plot in
Figure 4.7d where FCA is neglected. These results show that it is
more difficult to achieve self-pulsation in an APMR when signifi-
cant FCA is present.
Free carrier-induced self-pulsation in a microring resonator
can be explained in terms of the interactions between the optical
­resonance and the nonlinear dispersion caused by the TPA-generated
Nonlinear Optics Applications of Microring Resonators 211
(a)
0.07

Power in microring, P0 (W)


0.06
S SP
0.05
0.04
BS
0.03
0.02 δ=2
0.01 S τn = 2
0
0 0.5 1 1.5 2 2.5 3 3.5 4
Input power, Pin (mW)

(b) (c)
0.25 0.12

Power in microring, P0 (W)


Power in microring, P0 (W)

δ=2 τn = 2
SP 0.1 0.8
0.2 +
P0 , sp
11.2
1.4
– 0.08 1.6
0.15 P0, sp 1.6 SP
0
0.06
0.1 + 0
P0, bs τn,c 0.04 BS
BS
S 0
0.05 0.02 0
S –
P0 , bs
0 0
0 0.5 1 1.5 2 2.5 3 0 1 2 3 4 5
Normalized FC lifetime, τn Normalized frequency detune, δ

Figure 4.8 (a) Stability plots of a Si APMR showing the power in the
microring (P0) versus the input power (Pin) for δ = 2 and τn = 2. Black line
is the case where Kerr-induced index change and FCA are included; gray
line is the case where these effects are neglected. (b) Plot of the threshold
powers for bistability and self-pulsation as functions of the normalized
FC lifetime τn at a fixed δ = 2. (c) Stability map of microring power ver-
sus frequency detune showing regions of stability, bistability, and self-­
pulsation for a microring with τn = 2. Values of the contour lines in the
SP region are normalized periods of oscillations, Tsp /τ′L .

free carriers. More specifically, near a resonant frequency, the light


intensity buildup in the microring causes the FC density to increase,
which in turn causes a blue shift in the cavity mode due to FCD.
As a result, the light becomes detuned from the resonance, leading
to a decrease in the light intensity and hence the FC density in the
microring. However, as the FC density decreases, the blue shift in
the cavity mode also gets smaller, so the light becomes less detuned
from the resonance and begins to build up again in the cavity.
The resulting behavior is thus an oscillation of the light intensity in
the microring with a period comparable to the cavity lifetime.
212 Optical Microring Resonators

Self-pulsation can also arise from the competing interactions


between FC-induced index change and the photothermal refrac-
tion effect in a semiconductor microring resonator (Zhang et al.
2013). At moderate to high input powers, light absorption in the
microring can generate sufficient heat to induce a change in the
refractive index of the waveguide via the thermo-optic effect. In a
silicon waveguide, the index change due to photothermal refrac-
tion is positive, while the FC-induced index change is negative.
Thus FCD and photothermal refraction exert a kind of push-and-
pull effect on the resonant wavelength of a microring resonator
and, if the two processes have comparable time constants, can
result in oscillations of the field inside the microring. This self-pul-
sation behavior has been observed in silicon microring resonators
(Priem et al. 2005, Johnson et al. 2006, Pernice et al. 2010, Zhang
et al. 2014). However, due to the relatively long thermal time con-
stants in these devices, the oscillation frequencies are typically in
the MHz range or lower. On the other hand, self-pulsations caused
by the interaction of the ­nonlinear FCD effect with the resonance
mode, which are characterized by much faster oscillation frequen-
cies (in the GHz range), have not been observed, mainly due to
the deleterious effect of FCA and the short FC lifetime required to
observe this behavior.

4.3 All-Optical Switching in a


Nonlinear Microring Resonator
One of the applications of a material with an intensity-dependent
refractive index is all-optical switching, which involves the mod-
ification of an optical signal either by itself or by another opti-
cal signal. The former case, called self-switching, is based on the
process of SPM, whereby an optical signal switches itself from a
low-power state to a high-power state or vice versa. The latter case
typically employs a pump-and-probe switching scheme in which a
strong pump beam is used to modulate the amplitude of a weaker
probe signal through XPM. In this section, we consider both self-
switching and pump-and-probe switching in a nonlinear microring
resonator (Van et al. 2002a). In both schemes, a strong input pump
signal is used to cause an intensity-dependent shift of the microring
resonance, which translates into a modulation of the transmission
of either a probe signal or the pump itself.
Nonlinear Optics Applications of Microring Resonators 213

We begin in Section 4.3.1 by showing that under CW excitation,


the buildup of light intensity in the microring near a resonance can
lead to a substantial reduction in the nonlinear switching power
compared to a nonresonant device such as a Mach–Zehnder interfer-
ometer (MZI). In Section 4.3.2, we will treat the problem of nonlinear
pulse propagation in a microring resonator and show how SPM in the
resonator can lead to self-switching. Section 4.3.3 extends the analy-
sis to the case of pump-and-probe switching in a microring resonator
with intensity-dependent nonlinear refractive index. The effects of
FCD and FCA will also be examined for both switching schemes.

4.3.1 Enhanced nonlinear switching in a microring resonator


In a typical pump-and-probe switching configuration, a strong
pump signal is used to induce a nonlinear phase shift in a weaker
probe signal. This phase modulation can be converted into a change
in the transmitted amplitude of the probe signal through the use of
an interferometric device such as a MZI or a cavity such as a micror-
ing resonator. A comparison of the switching efficiency of these two
devices—one resonant and the other nonresonant—will serve to
highlight the advantage of microring resonators for nonlinear optics
applications. For simplicity, we will restrict our ­analysis below to
structures with instantaneous intensity-dependent r­ efractive index
under CW excitation.
We consider first a simple switch based on an unbalanced MZI
with phase difference δ = ϕ2 − ϕ1 between the two arms as shown in
Figure 4.9a. For a CW probe signal at frequency ωs applied to port 1,
its power transmission at port 4 is given by
2
sout  δ (4.185)
T= = cos 2   .
sin  2

Suppose an intense CW pump signal at frequency ωp is injected


into the upper arm of the MZI. The pump signal causes a non-
linear phase shift in the probe signal in the upper arm equal to
ϕNL = −2γPpksL, where γ is the nonlinear coefficient of the waveguide
given in Equation 4.51, Pp is the power of the pump beam, ks = ωs/c,
and L is the arm length of the MZI. The factor 2 appears because
the phase shift due to XPM is twice as large as that due to SPM.
The total phase difference of the two MZI arms is thus
δ = φ 2 − (φ1 − 2γPp ks L). (4.186)
214 Optical Microring Resonators
(a) (b)
1
Pump

Transmission, T
1 L 3 0.8 Maximum
Sin
φ1 0.6 switching slope
0.4
φ2
Sout 0.2
2 3 dB 3 dB 4
coupler coupler 0
0 0.2 0.4 0.6 0.8 1
Phase imbalance, δ/π

Figure 4.9 (a) Schematic of an unbalanced MZI for pump-and-probe


switching and (b) power transmission at port 4 as a function of the total
phase imbalance, δ = ϕ2 − (ϕ1 + ϕNL), between the two MZI arms.

Figure 4.9b plots the transmission of the probe signal with respect
to the total phase imbalance δ of the MZI. It is clear that the output
power of the probe signal, |sout|2, can be modulated by varying the
pump power. For example, a high input pump power will cause a
large nonlinear phase shift, which leads to a low transmitted probe
power. The switching efficiency of the MZI can be estimated from
the slope of the transmission curve,

dT dT dδ
= . (4.187)
dPp dδ dPp

The first derivative on the right-hand side (dT/dδ) can be evaluated


from Equation 4.185. Its maximum absolute value is 1/2, which
occurs when the MZI has a linear phase imbalance of ϕ2 − ϕ1 = π/2.
The second derivative can be computed from Equation 4.186 to give
dδ/dPp = 2γksL. The maximum switching slope of the MZI is thus

dT
= γks L, (4.188)
dPp max

which indicates that the switching efficiency increases linearly with


the MZI arm length L.
In an APMR the switching efficiency is enhanced by two sep-
arate effects. First, the effective nonlinear interaction length L is
increased by a factor roughly equal to the resonator finesse, ℱ ~ 1/κ 2,
where κ is the coupling coefficient. Second, the light intensity inside
the microring is amplified by a factor of FE2 ~ 1/κ2 at resonance (see
Equation 2.22). Thus we expect a reduction in the switching power
proportional to ~1/κ4 in an APMR compared to an MZI.
Nonlinear Optics Applications of Microring Resonators 215

To estimate the maximum switching power reduction achiev-


able in a microring resonator, we consider an APMR of radius
R, field coupling coefficient κ, and round-trip attenuation art.
We assume that the APMR is critically coupled so that a steep
switching slope can be obtained near a resonance. Under this
­condition, the linear ­transmission response of the device is given
by Equation 2.45 with τ = art,
2
s 2τ 2 (1 − cos φ)
Tap = out = , (4.189)
sin 1 + τ 4 − 2τ 2 cos φ

where ϕ is the microring round-trip phase. Suppose a weak CW


probe signal tuned near a microring resonance is applied to the
input port. The transmission of the probe signal is modulated by
applying a strong CW pump with power Pin,p to the input port, as
shown in Figure 4.10a. The pump beam is assumed to be of the
same polarization as the probe but tuned to a different resonant
frequency of the microring. Accounting for the nonlinear phase
shift induced by the pump beam through XPM, the total round-trip
phase of the probe signal in the microring is given by

φ = −(nr + 2γPr,p )ks L, (4.190)

where nr is the effective index of the waveguide, L = 2πR is the


microring circumference, and Pr,p = |Ap|2 is the pump power cir-
culating in the microring. Figure 4.10b plots the probe power trans-
mission Tap as a function of the round-trip phase detune Δϕ from

(a) (b)
Transmission slope, dTap/dφ

1 40
∆φmax
Power transmission, Tap

30
0.8
20

R 0.6 10
Ap
0
As 0.4 –10
Sin,p κ, τ Sout,p –20
0.2
Sin,s Sout,s –30
0 –40
–0.2 –0.15 –0.1 –0.05 0 0.05 0.1 0.15 0.2
Phase detune, ∆φ

Figure 4.10 (a) Schematic of an APMR for pump-and-probe switching


and (b) plots of the power transmission and transmission slope versus
phase detune of a critically coupled APMR with τ = art = 0.99.
216 Optical Microring Resonators

a microring resonance. Application of the pump beam causes the


phase detune to decrease (assuming γ > 0), so that the transmission
curve will effectively shift to the right in Figure 4.10b. Thus a probe
signal tuned to a point on the right shoulder of the transmission
curve will experience a decrease in the transmission.
Figure 4.10b also plots the variation in the slope of the transmis-
sion curve, which is given by

dTap 2τ 2κ 4 sin φ
= . (4.191)
dφ (1 + τ 4 − 2τ 2 cos φ)2

From the plot, we observe that the maximum transmission slope


occurs slightly off the resonance. The phase ϕmax at which the maxi-
mum slope occurs can be computed from the condition d2Tap/dϕ2 = 0.
Assuming that the microring has a very high Q factor (κ≪1), we get
cosϕmax ≈ 1 − κ4/6, which gives the maximum value for the slope of
the transmission curve as

dTap 3 3
≈ . (4.192)
dφ φmax
8κ 2

The switching efficiency of the APMR can be computed from


dTap dTap dφ dPr,p
= , (4.193)
dPin,p dφ dPr,p dPin,p

where the first derivative on the right-hand side is given by Equation


4.192. The second derivative is evaluated from Equation 4.190 to
give dϕ/dPr,p = −2γksL. The last derivative, dPr,p/dPin, gives the pump
power enhancement in the microring at ϕmax. From Equation 2.21 we
obtain the expression for the pump power in the APMR as follows:

κ 2 Pin,p
Pr,p = , (4.194)
1 + τ 4 − 2τ 2 cos φ

from which we evaluate the derivative at ϕmax to be dPr,p/dPin,p ≈ 3/4κ2.


Substituting the above results for the derivatives into Equation 4.193,
we obtain the maximum switching slope

dTap 9 3  γks L  4
≈   ∝ FE , (4.195)
dPin,p max
8  κ4 
Nonlinear Optics Applications of Microring Resonators 217

where FE ~ 1/κ is the field enhancement factor at resonance. The


above result shows that compared to an MZI having the same arm
length as the microring circumference, the microring resonator has
a switching slope that is proportional to the fourth power of the field
enhancement factor, or the square of the finesse F (since FE2 = ℱ/π).
For a microring resonator with power coupling ­efficiency κ2 ~ 1%,
the switching slope can be enhanced by as much as four orders of
magnitude.

4.3.2 Self-switching of a pulse in a
nonlinear microring resonator
In Section 4.2.1 we showed that under CW excitation, SPM effects
in a microring resonator with intensity-dependent refractive
index are enhanced by approximately the square of the field
enhancement factor (FE2) near a microring resonance. In this sec-
tion, we extend the analysis to a pulse propagating in a microring
resonator with an instantaneous intensity-dependent refractive
index. Simulation examples demonstrating self-switching in an
add-drop microring resonator will be given. We will also exam-
ine how free carrier effects can modify pulse propagation in the
microring.
We consider an ADMR with input and output field coupling
coefficients κ1 and κ2, respectively.* The microring is assumed
to have radius R, effective index nr, linear loss coefficient α0,
and nonlinear Kerr coefficient n2. An input signal of the form
sin (t) = sin (t)e jω0t, where sin(t) is the slowly varying pulse envelope,
is applied to the input port of the ADMR. The power-normalized
wave inside the microring can be expressed as
 ( z , t) = A( z , t)e j(ω0t −β0 z ) ,
A (4.196)

where β0 = nr(ω 0)ω 0/c. The coordinate z is defined along the cir-
cumference of the microring with z = 0 located just after the input
coupling junction, as shown in Figure 4.2a for an APMR device.
In the vicinity of a resonant frequency, the frequency dispersion
due to resonance dominates over the dispersion of the micror-
ing ­waveguide. We can thus neglect the effects of both phase
velocity dispersion and GVD in Equation 4.48 and write the

* The analysis in this section is also applicable to an APMR by setting κ2 = 0.


218 Optical Microring Resonators

equation for the propagation of the pulse envelope A(z, t) inside


the microring as

∂A α 0
+ A = − jγ |A|2 A, (4.197)
∂z 2

where γ is the nonlinear coefficient defined in Equation 4.51. The


solution of the above equation is
2
A( z , t) = A(0, t)e − α0 z/2 e − jγ|A( 0 ,t )| zeff , (4.198)

where the effective propagation distance zeff is given by Equation


4.57. At the input coupling junction of the ADMR, the field satisfies
the boundary condition

A(0, t) = τ1 A(L, t)e jφL − jκ 1sin (t), (4.199)

where L = 2πR is the microring circumference and ϕL = −β0L is the


linear round-trip phase. At the output coupling junction, the field
in the microring is attenuated by a factor τ 2 = 1 − κ 22 due to cou-
pling to the output waveguide. If the coupling is small, the power
removed from the microring at the output coupling junction may
be lumped into the linear waveguide loss so that the total round-
trip amplitude attenuation in the microring can be approximated
as τ2 art, where art = e −α0 L/2. Using Equation 4.198, we can write the
solution for A(L, t) as

A(L, t) = τ 2 art A(0, t)e jφNL , (4.200)

where ϕNL = −γ|A(0, t)|2Leff and Leff is the effective microring circum-
ference given by

 ( τ a )2 − 1 
Leff = 2πR  2 rt . (4.201)
 2 ln(τ 2 art ) 

Upon substituting Equation 4.200 into the boundary condition in


Equation 4.199 and solving for A(0, t), we get

− jκ 1sin (t)
A(0, t) = . (4.202)
1 − τ1τ 2 art e j( φL + φNL )
Nonlinear Optics Applications of Microring Resonators 219

Taking the absolute square of the above equation and defining the
input power Pin(t) = |sin(t)|2, we obtain the equation for the nonlin-
ear phase shift ϕNL,

{1 + F sin [(φ
2
L }
+ φ NL )/2] φ NL (t) = − γLeff FEmax
2
Pin (t), (4.203)

where F = 4τ1τ2 art/(1 − τ1τ2 art)2 is the contrast factor and


2
FEmax = κ 12 /(1 − τ1τ 2 art )2 is the maximum power enhancement in the
ADMR. Equation 4.203 can be solved using an iterative numerical
technique. Alternatively, for small nonlinear phase shifts, an approx-
imate solution can be obtained by linearizing the sine term around
ϕL (similar to Equation 4.147). At high input powers, multiple real
solutions for ϕNL may exist which indicate bistable (or multistable)
behavior. By substituting the solution for the nonlinear phase shift
into Equation 4.202, we obtain the complex envelope A(0, t) of the
signal in the microring. Finally, the output signals at the drop port
and through port of the ADMR are obtained from

sd (t) = − jκ 2 art A(0, t)e j( φL + φNL )/2 , (4.204)

st (t) = τ1sin (t) − jκ 1τ 2 art A(0, t)e j( φL + φNL ) . (4.205)

In Equation 4.204 we have made the approximation that the nonlin-


ear phase shift at the output coupling junction is equal to half the
value of the nonlinear round-trip phase.
Numerical results for a Gaussian pulse at the 1.55 μm wavelength
propagating in a nonlinear silicon ADMR are shown in Figure 4.11.
We assume that only Kerr nonlinearity is present in the Si waveguide
and neglect effects due to TPA and free carriers. The parameters
used for the Si waveguide are: effective index nr = 2.5, linear propa-
gation loss α0 = 2.5 dB/cm, Kerr coefficient n2 = 4.5 × 10−14 cm2/W,
and effective mode area Aeff = 0.15 μm2. The microring has radius
R = 5 μm and equal input and output power coupling coefficients
κ2 = 2%. The drop port transmission spectrum of the device has
a 3 dB bandwidth of 13 GHz. We apply an input Gaussian pulse
having power profile given by Pin(t) = P0exp[−(t/T0)2] with the cen-
ter frequency tuned to a microring resonance (ϕL = 2mπ). The pulse
width is set to be T0 = 50 ps, corresponding to a signal bandwidth
of 5.3 GHz, which is well below the 3 dB bandwidth of the micror-
ing so that signal distortion due to the filtering effect of the ADMR
is negligible. Figures 4.11a and b show the output powers at the
220 Optical Microring Resonators

(a) 1 (b) 1

0.8 0.8

Normalized power
Normalized power Pd/P0
0.01 × Pr/P0 0.1 × Pr/P0
0.6 0.6 Pt/P0

0.4 0.4 Pd/P0


0.2 0.2
Pt/P0
0 0
–3 –2 –1 0 1 2 3 –3 –2 –1 0 1 2 3
Time t/T0 Time t/T0

(c) (d)
1
0.6

Frequency chirp, δω/BW


Normalized peak power

0.8 P0 = 1 W
Pt/P0 0.4
Pd/P0
0.2
0.6
0
0.4 P0 = 0.01 W
–0.2
0.2 –0.4
–0.6
0
10–3 10–2 10–1 100 101 –3 –2 –1 0 1 2 3
Peak input power, P0 (W) Time t/T0

Figure 4.11 Powers at the drop port (Pd) and through port (Pt), and in the
microring (Pr) for (a) low peak input power, P0 = 0.01 W, and (b) high peak
input power, P0 = 1 W. The pulse width is fixed at T0 = 50 ps. (c) Switching
curves of the ADMR showing the dependence of the drop port trans-
mission (Pd/P0) and through port transmission (Pt/P0) on the peak input
power. (d) Frequency chirp of the signal in the microring for peak input
powers P0 of 0.01 W and 1 W, both with a pulse width of T0 = 50 ps.

drop port (Pd) and through port (Pt) of the microring for peak input
pulse power P0 = 0.01 W and 1 W, respectively. These peak power
levels correspond to pulse energies of 0.9 pJ and 90 pJ, respectively.
The power in the microring (Pr) is also shown for each case. At low
input power (Figure 4.11a), the input signal is nearly in resonance
with the microring, so most of the pulse power appears at the drop
port and very little power appears at the through port. When the
input power is increased (Figure 4.11b), the signal becomes detuned
with respect to the microring resonance, causing most of the pulse
power to be switched to the through port. We note that the peak
power in the microring is enhanced by a factor of about 45 for the
low input power case and a factor of 7 for the high input power
case. This enhancement of the microring power helps amplify the
Nonlinear Optics Applications of Microring Resonators 221

nonlinear effects in the resonator and lower the switching power.


Figure 4.11c shows the variations of the peak power transmissions
at the drop port (Pd/P0) and the through port (Pt/P0) with respect to
the input power. From the plot, we find that the maximum switch-
ing slope occurs around a peak input power of about 100 mW.
It is also instructive to see how the frequency spectrum of the
signal inside the microring is modified by the Kerr nonlinearity.
We found in Section 4.1.3 that a pulse propagating in a waveguide
with instantaneous Kerr nonlinearity experiences spectral broad-
ening caused by frequency chirping. Due to resonance enhance-
ment of SPM effects, the spectral broadening experienced by a
pulse in a microring resonator can be much larger than in a straight
waveguide. Figure 4.11d shows the frequency chirp experienced by
a Gaussian pulse propagating in the silicon ADMR for two different
values of the peak input pulse power: P0 = 0.01 W and 1 W. We see
that the frequency spectrum of the signal broadens by around 8%
of the bandwidth at the low input power and as much as 60% at the
high input power.
The switching speed of a microring resonator with instantaneous
nonlinearity is limited by the cavity lifetime. In a semiconductor
microring resonator in which substantial TPA also occurs, FCA and
dispersion can greatly modify the device response, especially at high
input powers and large input pulse widths. In ­particular, if the FC
lifetime is much longer than the cavity lifetime, the switching speed
of the device will be mainly limited by the former time ­constant. To
take into account the effects of free ­carriers on the switching response
of an ADMR, the nonlinear wave equation (4.68) must be solved along
with Equation 4.80 for a pulse propagating in the microring subject
to the boundary condition at the input coupling junction given by
Equation 4.199. To ­simplify the ­solution, we may model the output
coupler of the ADMR as a source of loss and lump it with the linear
propagation loss of the microring waveguide, so that the effective
linear loss coefficient becomes α0,eff = α0 − ln(τ2)/πR.
The effects of TPA and free carriers on the propagation of a
Gaussian pulse in a Si ADMR are illustrated by the plots in Figure
4.12. The microring parameters are the same as those in the previ-
ous example. In addition, we also assume the following FC param-
eters for silicon around the 1.55 μm wavelength: TPA coefficient
α2 = 0.8 cm/GW, FC refraction volume σr = −5.3 × 10−21 cm3, FCA
cross section σa = 1.45 × 10−17 cm2, and carrier lifetime τfc = 1 ns.
Figure 4.12a plots the transmitted powers at the drop port (Pd) and
222 Optical Microring Resonators

(a) 1 (b) 30

Normalized power 0.8 25

Normalized power
20
0.6
Pt/P0
15
0.4
Pd/P0 10 Pr/P0
0.2 5
0 0
–3 –2 –1 0 1 2 3 –3 –2 –1 0 1 2 3
Time t/T0 Time t/T0

(c) ×1016 (d)


8 1.5
FC density, Nfc(t) (cm–3)

Frequency chirp, δω/BW


7
Kerr + FC
6 1
5
4 0.5 Kerr only
3
2 0
1
0 –0.5
–4 –3 –2 –1 0 1 2 3 4 –3 –2 –1 0 1 2 3
Time t/T0 Time t/T0

Figure 4.12 Nonlinear response of a Si ADMR to a Gaussian input pulse


with peak power P0 = 0.1 W and pulse width T0 = 50 ps: (a) transmitted
powers at the drop port (Pd) and through port (Pt); (b) power in the micror-
ing (Pr) when both Kerr nonlinearity and FC are present (solid lines) and
when only Kerr nonlinearity is present (dotted lines). (c) Time evolution of
the FC density (Nfc) inside the microring. (d) Frequency chirp of the pulse
in the microring when both Kerr and FC nonlinear effects are present
(solid line) and when only Kerr nonlinearity is present (dotted line).

through port (Pt) of the ADMR for an input Gaussian pulse with
peak power P0 = 0.1 W and pulse width T0 = 50 ps. The pulse power
and FC density inside the microring are shown in Figures 4.12b
and c, respectively. For comparison, we also show in Figures 4.12a
and b the pulse signals (dotted lines) when FC effects are neglected
and only Kerr nonlinearity is present in the microring. It is appar-
ent that FC effects significantly distort and attenuate the pulse
inside the microring and also the output signal at the drop port.
Figure 4.12d compares the frequency chirp of the pulse inside the
microring with and without FC effects. In the presence of free
­carriers, we observe significant positive chirp at the leading edge of
the pulse due to a rapid increase in the generated FC density.
Nonlinear Optics Applications of Microring Resonators 223

4.3.3 Pump-and-probe switching in a
nonlinear microring resonator
In a typical all-optical switching application, a control (pump) wave
is used to modulate the amplitude of a signal (probe) wave. Such
a pump-and-probe switching operation can be accomplished in a
microring resonator through the process of XPM, whereby a strong
pump wave is used to induce a nonlinear phase shift in the probe
signal, causing a modulation of its transmitted amplitude.
To analyze the interaction of the pump and probe waves in a
microring resonator with an instantaneous intensity-dependent
refractive index, we consider an ADMR with input pulses for both
the pump and probe of the form

sin,1 (t) = sin,1 (t)e jω1t , sin,2 (t) = sin,2 (t)e jω 2t , (4.206)

where sin,1(t) and sin,2(t) are the envelopes of the pump pulse and
probe pulse, respectively, and the frequencies ω1 and ω2 are tuned
to two different resonances of the microring. For simplicity, we
assume that the pump and probe waves have the same polarization,
and the peak power of the pump pulse is much larger than that of
the probe pulse so that we can neglect XPM of the pump by the
probe. If we also neglect both phase velocity dispersion and GVD,
then the equation governing the pump pulse envelope A1(z, t) in the
microring and its solution are given by Equations 4.197 and 4.198,
respectively. The equation for the probe pulse envelope A2(z, t) in
the microring is

∂A2 α 0 2
+ A2 = − j 2γ 2 A1 ( z , t) A2 ( z , t), (4.207)
∂z 2

where γ 2 = nc2 n2ω 2 /neff


2
(ω 2 )cAeff (ω 2 ) and the factor 2 on the right-
hand side appears due to the fact that XPM effect is twice as
strong as SPM. We have also assumed in Equation 4.207 that
the probe signal experiences negligible SPM and has the same
linear ­
­ propagation loss as the pump wave. The solution for
A 2(z, t) is
2
A2 ( z , t) = A2 (0, t)e − α0 z/2 e − j 2 γ 2|A1 ( 0 ,t )| zeff , (4.208)

where zeff is the effective propagation distance given by Equation


4.57.
224 Optical Microring Resonators

At the input coupling junction of the ADMR, the probe signal


must satisfy the boundary condition

A2 (0, t) = τ1 A2 (L, t)e jφL , 2 − jκ 1sin , 2 (t), (4.209)

where κ1 and τ1 are the input coupling and transmission coeffi-


cients, respectively, L = 2πR, and ϕL,2 = −β0(ω2)L is the linear round-
trip phase of the probe beam. Substituting the solution for A2(L, t) in
Equation 4.208 into the above boundary condition and solving for
A2(0, t), we obtain

− jκ 1sin,2 (t)
A2 (0, t) = , (4.210)
1 − τ1τ 2 art e j( φL,2 + φNL,2 )

where art = e −α0 L/2, ϕNL,2 = −2γ2|A1(0, t)|2Leff, and Leff is the effective
microring circumference. From the solutions for A1(0, t) and A2(0, t),
we can obtain the envelopes of the transmitted pump and probe
pulses at the drop port and through port of the ADMR by applying
Equations 4.204 and 4.205 to each of the pump and probe waves.
Figure 4.13 shows a simulation example of pump-and-probe
switching in a Si ADMR with instantaneous Kerr nonlinearity.
The linear parameters and Kerr coefficient of the microring are the
same as those in the example of Figure 4.11. Free carrier effects are
neglected. The input pump beam is a Gaussian pulse with peak
power P0 = 1 W, pulse width T0 = 50 ps, and center frequency tuned
to a microring resonance near the 1.55 μm ­wavelength. The probe
beam is a CW signal with 1 μW average power whose frequency
is tuned to a different but nearby resonance of the microring.
Both the pump and probe beams are assumed to be TE polarized.
The pump pulse inside the microring and those appearing at the
drop port and through port of the ADMR are the same as shown
in Figure 4.11b. Figure 4.13a shows the nonlinear phase change of
the probe signal induced by the pump pulse inside the microring
resonator. This phase change causes the probe signal to become
detuned from the microring resonance. As a result, the probe
power is discharged from the microring during the duration of the
pump pulse, as seen in Figure 4.13b, and only returns to its reso-
nant state after the pump pulse has passed. At the through port of
the ADMR, the probe signal appears as a pulse, while at the drop
port, it has the same inverted pulse shape as the probe signal inside
the microring.
Nonlinear Optics Applications of Microring Resonators 225
(a) (b)
0.5 1

Nonlinear phase, |φNL|/π 0.4 Probe 0.8 Pd,2/P0 Pt,2/P0

Normalized power
0.3 Pump 0.6

0.2 0.4 0.01 × Pr,2/P0

0.1 0.2

0 0
–4 –3 –2 –1 0 1 2 3 4 –4 –3 –2 –1 0 1 2 3 4
Time t/T0 Time t/T0

Figure 4.13 Pump-and-probe switching in a Si ADMR with instanta-


neous Kerr nonlinearity and negligible FC effects. The input pump beam
is a Gaussian pulse with peak power P0 = 1 W and pulse width T0 = 50 ps.
The probe beam is a CW signal with 1 μW average power. (a) Nonlinear
phase changes experienced by the pump and probe signals in the micror-
ing and (b) probe signal powers at the drop port (Pd,2 = |sd,2|2) and through
port (Pt,2 = |st,2|2), and inside the microring (Pr,2 = |A2(0,t)|2).

The presence of free carriers in the microring resonator can have


a significant impact on the pump and probe pulses in the resona-
tor. To simplify the analysis, we consider only the effects of TPA on
the pump beam and neglect TPA of the probe beam since the peak
pump power is much larger than that of the probe. Accounting for
both Kerr and FC-induced nonlinearities, we first solve Equations
4.68 and 4.80 for the pump pulse envelope A1(z, t) and the average
FC density N fc ( z , t) in the microring. The equation for the probe
beam is

∂A2 α 0
+ A2 = − j 2γ K,1 |A1 ( z , t)|2 A2 ( z , t) − j 2γ fc,1 N fc A2 ( z , t), (4.211)
∂z 2 
where γK,1 and γfc,1 are the nonlinear Kerr and FC coefficients defined
in Equations 4.69 and 4.70, respectively, for the pump beam.
As a numerical example, we show in Figures 4.14a and b the
pump and probe pulses, respectively, at the drop port, through port,
and inside a Si ADMR. The parameters of the device are the same as
those in the example of Figure 4.12. The input pump pulse has peak
power P0 = 2.2 mW, pulse width T0 = 50 ps, and c­enter frequency
tuned to a microring resonance. The probe signal is a CW signal with
1 μW average power tuned to a nearby resonance. Due to the strong
FC-induced nonlinearity in the microring, the probe signal can be
226 Optical Microring Resonators

(a) 1 (b) 1
Pd,1/P0 Pd,2/P0 Pt,2/P0
0.8 0.8

Normalized power
Normalized power
0.6 0.01 × Pr,1/P 0.6

0.4 10 × Pt,1/P0
0.4

0.2 0.2 0.01 × Pr,2/P0

0 0
–3 –2 –1 0 1 2 3 –3 –2 –1 0 1 2 3
Time t/T0 Time t/T0

Figure 4.14 Pump-and-probe switching in a Si ADMR: plots of the


­powers at the drop port and through port, and in the microring for
(a) the pump beam and (b) the probe beam. The input pump pulse has
peak power P0 = 2.2 mW and pulse width T0 = 50 ps. The probe beam is a
­continuous wave with 1 μW average power.

switched with a much lower pump pulse power compared to the


example in Figure 4.13, where only Kerr nonlinearity is present.
Experimental demonstration of pump-and-probe switching in
a microring resonator was first reported in Van et al. (2002b) using
free carrier effects in a vertically coupled GaAs/AlGaAs ADMR.
On and off switching times of less than 100 ps were obtained,
which were limited by the FC lifetime. It was also suggested that
the switching speed of the device could be improved by applying a
DC bias to sweep out the carriers. Other demonstrations of switch-
ing based on FC were also reported in a laterally coupled GaAs/
AlGaAs microracetrack (Ibrahim et al. 2003a), as well as in a silicon
microring resonator (Almeida et al. 2004). XPM was also employed
to demonstrate several photonic logic gate operations with micror-
ings, including AND/NAND gates based on InP and GaAs APMRs
(Ibrahim et al. 2003b), and a NOR gate based on two cascaded GaAs/
AlGaAs microring resonators (Ibrahim et al. 2004).

4.4 FWM in a Nonlinear Microring Resonator


Frequency generation based on FWM in a microring resonator was
first theoretically studied in Van and Little (1999) and experimen-
tally demonstrated in a GaAs/AlGaAs microring resonator in Absil
et al. (2000). Due to the field enhancement near a microring reso-
nance, much higher frequency conversion efficiency can be achieved
in the resonator than in a straight waveguide with the same length
Nonlinear Optics Applications of Microring Resonators 227

as the microring circumference. Applications of frequency genera-


tion in microring resonators have since been extended to frequency
comb generation (Del’Haye et al. 2008), hyperparametric oscillation
(Razzari et al. 2010), and the creation of entangled photon pairs by
spontaneous FWM (Clemmen et al. 2009).
In this section, we develop a formalism based on power coupling
in space to analyze degenerate FWM in an APMR with instanta-
neous Kerr nonlinearity. We consider an APMR with radius R, field
coupling coefficient κ and linear propagation loss α0 in the micror-
ing waveguide. A pump wave with power Pin,p and frequency ωp
and a signal wave with power Pin,s and frequency ωs are applied
to the input waveguide, as depicted in Figure 4.15. For simplicity
we assume that both the pump and signal waves have the same
linear polarization. The pump and signal frequencies ωp and ωs
are tuned to two different resonances of the microring, with the
separation between them denoted by Δω = ωs − ωp. FWM generates
an idler wave at frequency ωi = 2ωp − ωs = ωp − Δω. If the separation
Δω between the pump and signal frequencies is small (typically
equal to one FSR of the microring), the dispersion in the microring
­resonances is small so that the idler frequency will also coincide
to a microring resonant frequency. We will examine the effects of
resonance detuning and phase mismatch caused by dispersion in
the microring in more detail at the end of the section.
Under the assumption of no pump depletion, we can solve the
nonlinear wave equation for the pump beam and the coupled wave
equations for the signal and idler waves separately. Defining z as
the propagation distance along the circumference of the microring
as shown in Figure 4.15, we can write the equation for the pump
wave Ap(z) in the microring resonator as

dAp α 0
+ Ap = − jγ |Ap|2 Ap , (4.212)
dz 2

Ap, As, Ai
z= 0
Pin,p, Pin,s Pout,p, Pout,s, Pout,i
κ, τ

Figure 4.15 Degenerate FWM in a nonlinear APMR.


228 Optical Microring Resonators

where γ = γ(ωp) is the nonlinear coefficient in Equation 4.51.


The ­solution of the above equation is given by Equation 4.142,
− jγPp zeff
Ap ( z) = Pp e − α0 z/2 e , (4.213)

where Pp = |Ap(0)|2 is the pump power in the microring at z = 0 and


zeff = (1 − e − α0 z )/α 0 is the effective propagation distance. At the cou-
pling junction between the microring and the straight waveguide,
the pump wave satisfies the boundary condition
jφL,p
Ap (0) = τAp (L)e − jκ Pin,p , (4.214)

where τ = 1 − κ 2 , L = 2πR, and ϕL,p = −nr(ωp)ωpL/c is the linear


round-trip phase. By substituting the solution in Equation 4.213
into the above boundary condition, we obtain an equation similar
to Equation 4.145 whose solution gives the pump power Pp in the
microring.
For the signal and idler waves, the coupled wave equations are
given by Equations 4.129 and 4.130,

dAs α 0
dz
+
2
(
As = − jγ 2|Ap|2 As + Ap2 Ai∗ e j∆βz , ) (4.215)

dAi∗ α 0 ∗
dz
+
2
(
Ai = jγ 2|Ap|2 Ai∗ + ( Ap∗ )2 As e − j∆βz , ) (4.216)

where Δβ = βs + βi − 2βp is the wave vector mismatch. For simplic-


ity we have neglected dispersion in both the linear loss and the
nonlinear coefficient so that α0(ωs) ≈ α0(ωi) ≈ α0 and γ(ωs) ≈ γ(ωi) ≈ γ.
Writing the solutions for the signal and idler waves as
− j 2 γPp zeff
As ( z) = Bs ( z)e − α0 z/2 e , (4.217)

j 2 γPp zeff
Ai∗ ( z) = Bi∗ ( z)e − α0 z/2 e , (4.218)

and making use of the pump wave solution in Equation 4.213, we


can simplify Equations 4.215 and 4.216 to

dBs
= − jγPp e − α0 z Bi∗ e jψ ( z ) , (4.219)
dz
Nonlinear Optics Applications of Microring Resonators 229

dBi∗
= jγPp e − α0 z Bs e − jψ ( z ) , (4.220)
dz
where ψ(z) = Δβz + 2γPpzeff. In general, the coupled wave equations
in Equations 4.219 and 4.220 must be solved numerically. However,
if we neglect the effect of attenuation of the pump beam due to lin-
ear loss in the microring, we can set e − α0 z ≈ 1 and zeff ≈ z, in which
case Bs and Bi will have solutions of the form similar to Equations
4.132 and 4.133. We can thus write the solutions for As and Ai∗ as

As ( z) = [as cosh( gz) + bs sinh( gz)]e − α0 z/2 e − jΩ( z ) , (4.221)

Ai∗ ( z) = [ai∗ cosh( gz) + bi∗ sinh( gz)]e − α0 z/2 e jΩ( z ) , (4.222)

where

g = ( γPp )2 − (K/2)2 , (4.223)

Ω( z) = 2γPp zeff − Kz/2, (4.224)

K = ∆β + 2γPp . (4.225)

At the coupling junction, the signal and idler waves satisfy the
boundary conditions

As (0) = τAs (L)e jφL,s − jκ Pin,s , (4.226)

Ai∗ (0) = τAi∗ (L)e − jφL,i , (4.227)

where ϕL,k = −nr(ωk)ωkL/c is the linear round-trip phase at fre-


quency ωk = {ωs, ωi}. In addition, by evaluating the coupled wave
equations (4.215) and (4.216) at z = 0, we also have the boundary
conditions

dAs α0 α
=− As (0) − jγPp [2 As (0) + Ai∗ (0)] = gbs − 0 as
dz z=0 2 2
− j( γPp − ∆β/2)as , (4.228)

dAi∗ α0 ∗ α
=− Ai (0) + jγPp [2 Ai∗ (0) + As (0)] = gbi∗ − 0 ai∗
dz z=0
2 2
+ j( γPp − ∆β/2)ai∗ . (4.229)
230 Optical Microring Resonators

Applying the boundary conditions in Equations 4.226 through 4.229


to the solutions for As and Ai∗ in Equations 4.221 and 4.222, we solve
for the coefficients as, bs, ai, and bi to get

1 − τart cosh( gL)e − jϕi ∗


bi∗ = ai = f1 ai∗ , (4.230)
τart sinh( gL)e − jϕi

jγPp
ai∗ = as = f 2 as , (4.231)
gf1 − jK/2

bs = − j( γPp f 2 + K/2)as /g = f 3 as , (4.232)

− jκ Pin,s
as = = f 4 Pin,s , (4.233)
1 − τart [cosh( gL) + f 3 sinh( gL)]e jϕs

where φs = ϕL,s − Ω(L), φi = ϕL,i − Ω(L), and art = e −α0 L/2 is the round-
trip amplitude attenuation due to linear loss in the microring.
Similar to the case of FWM in a straight waveguide, we find
from Equation 4.223 that in order to have gain in the microring
waveguide, we require (γPp)2 − (K/2)2 > 0, or −4γPp < Δβ < 0. The
maximum gain is gmax = γPp, which occurs when K = 0 or Δβ = −2γPp.
Since the pump power inside the microring is enhanced by approx-
imately a factor of FE2 ~ 1/κ2 with respect to the input pump power,
the range of wave vector mismatch for which gain is achieved
becomes −4γPin,p/κ2 < Δβ < 0, which is broadened by a factor of
FE2 compared to FWM in a straight waveguide. However, since
the circumference of a microring resonator is typically short, the
phase mismatch per round-trip, ΔβL, is small so that wave vector
mismatch is less important for FWM in a microring resonator than
in a straight waveguide. In fact, for values of Δβ outside the range
[−4γPp,0] the idler wave still experiences gain per round-trip even
though the waveguide gain g is imaginary.*

* For Δβ > 0 or Δβ < −4γPp, the gain is g = jg′ where g ′ = (K/2)2 − ( γPp )2 . The solu-
tions for the signal and idler waves are
As ( z) = [as cos( g ′z) + bs sin( g ′z)]e − α0z/2 e − jΩ( z ) ,

Ai∗ ( z) = [ai∗ cos( g ′z) + bi∗ sin( g ′z)]e − α0z/2 e jΩ( z ) .

In this case both the signal and idler waves can still exhibit gain for small propa-
gation distance z.
Nonlinear Optics Applications of Microring Resonators 231

The frequency conversion efficiency of the APMR can be defined


as the ratio of the idler wave power at the output port to the signal
wave power at the input port, ηr = Pout,i/Pin,s. The idler wave power
at the output port is

Pout,i = κ 2 |Ai (L)|2 = (κart )2 |ai||cosh(


2
gL) + f1 sinh( gL)|2 = (κ/τ)2 |ai |2 .
(4.234)

The input signal power is related to the signal wave in the micror-
ing by Pin,s = |as/f4|2. Thus the conversion efficiency of the APMR is

ηr = (κ/τ)2 | f 2 ||
2
f 4 |2 . (4.235)

In the above expression, |f2|2 is the ratio of the idler power to sig-
nal power in the microring (i.e., the conversion efficiency in the
microring), and |f4|2 gives the power enhancement of the signal
wave due to resonance. Under the condition of maximum gain in
the microring waveguide (K = 0 and gmax = γPp), and assuming that
the signal and idler waves are tuned to coincide with the micror-
ing resonances so that φs = 2msπ and φi = 2miπ, we can evaluate the
­conversion efficiency of the APMR to get

κ 4 art2 sinh 2 ( g max L)


ηr = . (4.236)
( )( )
2 2
1 − τart e gmax L 1 − τart e − gmax L

The above expression shows that when the gain gmax is such that
e gmax L = 1/τart , that is, the round-trip gain completely compen-
sates for the round-trip loss, the conversion efficiency is infinite.
However, we have neglected the effect of pump depletion in our
analysis, which becomes important at high conversion efficiencies.
In practice, pump depletion and resonance detuning of the pump,
signal, and idler waves due to SPM and XPM limit the maximum
achievable conversion efficiency in the microring.
Since the product gmaxL is typically small, we can make the
approximation e gmax L ≈ e − gmax L ≈ 1. Further assuming that the micror-
ing resonator has a high Q factor (τ ≈ 1 and art ≈ 1), we can estimate
the conversion efficiency as follows:

κ 4 sinh 2 ( g max L) κ 4 (1 + τ)4 sinh 2 ( g max L)  16 


ηr ≈ = ≈  4  ηwg , (4.237)
(1 − τ)4 (1 − τ 2 )4 κ 
232 Optical Microring Resonators

where ηwg = sinh2(gmaxL) is the maximum conversion efficiency in a


straight waveguide with the same length as the microring circum-
ference. Since the field enhancement in the microring is FE ~ 1/κ,
the above result shows that the conversion efficiency in an APMR
is enhanced by roughly the fourth power of the field enhancement
factor.
As a numerical example, we show in Figure 4.16 the FWM con-
version efficiency in a Si APMR around the 1.55 μm ­wavelength and
its dependence on various parameters. The microring is assumed to
have radius R = 20 μm, coupling coefficient κ = 0.4, effective index
nr = 2.5, and linear loss coefficient α0 = 2.5 dB/cm. The Kerr coeffi-
cient of silicon is n2 = 4.5 × 10−14 cm2/W and the effective mode area
of the waveguide is Aeff = 0.15 μm2. Figure 4.16a plots the conversion
efficiency as a function of the wave vector m ­ ismatch Δβ for several

(a) (b)
0.32 101
Pin,p = 250 mW
Conversion efficiency, ηr

Conversion efficiency, ηr

0.3 Pin,p = 100 mW 100 100 mW


0.28 50 mW
0.26 250 mW 10–1

0.24 10–2
50 mW
0.22
10–3
0.2
0.18 10–4
–20 –15 –10 –5 0 5 10 15 20 –10 –8 –6 –4 –2 0 2 4 6 8 10
Wave vector mismatch, ∆β/γPp Signal phase detune, ∆φL,s/γPpL

(c) (d)
102 0.35
Conversion efficiency, ηr

Conversion efficiency, ηr

0.3 Pin =
101 |f4|2 250 mW
0.25 100 mW
100
ηr 0.2 50 mW
10–1
0.15
10–2 |f2|2 0.1
10–3 0.05
10–4 –3 0
10 10–2 10–1 100 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Input pump power, Pin,p (W) Coupling coefficient, κ

Figure 4.16 Frequency conversion efficiency (ηr) in a Si APMR: (a) ηr ver-


sus wave vector mismatch Δβ for ΔϕL,p = ΔϕL,s = ΔϕL,i = 0, (b) ηr versus sig-
nal phase detune ΔϕL,s for Δβ = 0, (c) ηr versus input pump power Pin,p for
Δβ = 0 and ΔϕL,p = ΔϕL,s = ΔϕL,i = 0, (d) ηr versus coupling coefficient κ for
Δβ = 0 and ΔϕL,p = ΔϕL,s = ΔϕL,i = 0.
Nonlinear Optics Applications of Microring Resonators 233

input pump powers. The linear phase detunes of the pump, sig-
nal, and idler waves are assumed to be zero (ΔϕL,p = ΔϕL,s = ΔϕL,i = 0).
We observe that the conversion efficiency varies only slightly with
the wave vector mismatch because of the short round-trip length
(circumference) of the microring. Also, the maximum conversion
efficiency does not occur at Δβ = −2γPp as for FWM in a straight
waveguide because XPM causes a slight detuning of the signal and
idler waves from the microring resonances. This resonance detun-
ing also causes a drop in the cosnversion efficiency as the input
pump power is increased from 100 mW to 250 mW.
Figure 4.16b plots the conversion efficiency versus the linear
phase detune of the signal wave (ΔϕL,s) from the resonance for
several input pump powers under the condition of zero wave vec-
tor mismatch (Δβ = 0). We assume that the idler wave also has the
same linear phase detune as the signal wave (ΔϕL,i = ΔϕL,s), whereas
the pump wave has zero phase detune (ΔϕL,p = 0). The plot shows
that the conversion efficiency depends strongly on the linear
phase detune of the signal, reaching a peak value at ΔϕL,s = 2γPpL.
At this linear phase detune, both the conversion efficiency inside
the microring (|f2|2) and the power enhancement of the signal wave
(|f4|2) are maximized, resulting in maximum idler wave power at
the output of the APMR. Note that although the plot shows that the
conversion efficiency exceeds 100% at high pump powers, in reality
the conversion efficiency is limited by pump depletion.
Figure 4.16c shows the dependence of the conversion efficiency
on the input pump power under the condition of zero wave vector
mismatch (Δβ = 0) and zero phase detunes (ΔϕL,p = ΔϕL,s = ΔϕL,i = 0).
Also shown are the variations of the conversion efficiency inside
the microring (|f2|2) and the power enhancement of the signal wave
(|f4|2) with the input pump power. We observe that for low pump
powers, the conversion efficiency increases with the pump power
due to increase in gain in the microring, reaching a maximum for
Pin,p ~ 0.1 W. Above this pump power, the signal and idler waves
become increasingly detuned from the microring resonances, as
evident from the decrease in |f4|2, causing the conversion efficiency
to drop as the pump power is further increased.
Figure 4.16d shows the dependence of the conversion efficiency
on the coupling coefficient of the APMR. We observe that the con-
version efficiency initially increases as κ is increased, reaching a
peak value before decreasing again for larger κ values. This varia-
tion generally follows the variations of the enhancements of the
234 Optical Microring Resonators

pump, signal, and idler powers inside the microring with respect to
the coupling coefficient.
We now examine the effect of waveguide dispersion on the
resonance detuning of the idler wave and the phase-matching
­
­condition inside the microring. We assume that the pump beam
and the signal wave are tuned exactly to two different resonant
­frequencies, Ωp and Ωs, respectively, of the microring:
mp c
ω p = Ωp = , (4.238)
nr (Ω p ) R

ms c
ω s = Ωs = = ω p + ∆ω , (4.239)
nr (Ωs ) R

where nr(ω) is the effective index of the microring waveguide, mp


and ms are the resonance mode numbers, and Δω = ωs − ωp is the
frequency separation between the pump and the signal wave. The
idler wave is generated at frequency ωi = 2ωp − ωs, which is detuned
from the resonance mode Ωi of the microring by

mi c
δω i = ω i − Ωi = (2ω p − ω s ) − , (4.240)
nr (Ωi ) R

where mi = 2mp − ms. Writing the above equation as

 mp c nr (Ω p )   ms c nr (Ωs ) 
δω i = 2 ω p −  − ω s − , (4.241)
 nr (Ω p )R nr (Ωi )   nr (Ωs )R nr (Ωi ) 

we see that if there is no waveguide dispersion, nr(Ωp) = nr(Ωs) =


nr(Ωi) = nr, δωi is zero, and the idler wave coincides exactly with
the resonance mode mi of the microring (ωi = Ωi). However, due
to ­dispersion, the idler wave will be slightly detuned from the
microring resonance Ωi. To determine the amount of detuning,
we approximate the effective index of the microring waveguide
in terms of a first-order Taylor series expansion around the pump
beam frequency,

nr (ω) = nr (Ω p ) + (ω − Ω p )D1 , (4.242)

where D1 = dnr/dω evaluated at ω = Ωp. The derivative D1 can be


related to the group index of the microring waveguide through the
Nonlinear Optics Applications of Microring Resonators 235

formula ng = nr(Ωp) + ΩpD1. Using the above Taylor series approxi-


mation, we can express the index ratios nr(Ωp)/nr(Ωi) and nr(Ωs)/nr(Ωi)
in Equation 4.241 as
nr (Ω p ) nr (Ω p )
= ≈ 1 − (Ωi − Ω p )D1 /nr , (4.243)
nr (Ωi ) nr (Ω p ) + (Ωi − Ω p )D1

nr (Ωs ) nr (Ω p ) + (Ωs − Ω p )D1


= ≈ 1 + (Ωs − Ωi )D1 /nr , (4.244)
nr (Ωi ) nr (Ω p ) + (Ωi − Ω p )D1

where nr = nr(Ωp). With the above approximations, Equation 4.241


can be simplified to

δω i ≈  2Ω p (Ωi − Ω p ) + Ωs (Ωs − Ωi ) D1 /nr . (4.245)

Substituting Ωs = Ωp + Δω and Ωi = Ωp − Δω − δωi into Equation


4.245 and solving for δωi, we obtain the following expression for
the resonance detune (or frequency walk-off) of the idler wave:

2D1∆ω 2
δω i ≈ . (4.246)
ng − D1∆ω

The above equation shows that the amount of resonance detun-


ing experienced by the idler wave increases as the square of the
frequency separation Δω between the pump and signal waves.
As the idler wave is detuned from the microring resonance, the con-
version efficiency decreases since the wave is no longer enhanced
by resonance. Resonance detuning due to waveguide dispersion
­ultimately limits the maximum achievable frequency conversion
bandwidth, or gain bandwidth, of FWM in a microring resonator.
Next we examine how the phase-matching condition is affected
by dispersion in the microring waveguide. The wave vector
­mismatch Δβ is given by

∆β = β(ω s ) + β(ω i ) − 2β(ω p ). (4.247)

We can express the propagation constant in terms of a Taylor series


around the pump frequency ωp as (Marhic et al. 1996)

βn
β(ω) = β(ω p ) + ∑ n! (ω − ω ) ,
n =1
p
n
(4.248)
236 Optical Microring Resonators

where β n = (∂ β/∂ω )ω =ω p. Upon substituting Equation 4.248 into


n n

4.247, we find that all the odd-order terms cancel out and only the
even-order terms contribute to the wave vector mismatch:

1
∆β = ∆ω 2β 2 + ∆ω 4β 4 + . (4.249)
12

Thus to the lowest order, the wave vector mismatch also increases as
the square of the frequency separation Δω between the pump and
signal waves. However, since the round-trip length of a microring
resonator is typically short, the total round-trip phase mismatch
ΔβL caused by dispersion is small so that phase ­mismatch plays
a less critical role than resonance detuning in ­limiting the FWM
­conversion bandwidth of an APMR.
Finally we note that, in a semiconductor microring resonator
with significant TPA, FCD and FCA can also have deleterious effects
on the FWM conversion efficiency and bandwidth (Lin et al. 2007).
It was reported in Ong et al. (2013) that by actively removing the
generated carriers using a reverse-biased pin junction, a twofold
improvement in the FWM c­ onversion efficiency in a silicon micror-
ing resonator could be achieved.

4.5 Summary
This chapter reviews important nonlinear optics applications
based on an intensity-dependent nonlinear refractive index in a
microring resonator. These applications include bistable devices,
self-­oscillations, all-optical switching, and frequency conversion.
In many of these applications, the large field enhancement near a
microring resonance helps amplify the nonlinear effects, leading
to a reduction in the required input power and hence better device
efficiency compared to similar processes in a straight w ­ aveguide.
For each application, a formalism based on power coupling in
space is also developed to provide a semi-analytical analysis of
the nonlinear optical process involved. While only intensity-
dependent nonlinear refraction and absorption are considered
in this chapter, the formalisms developed can also be applied to
analyze microring devices with other types of nonlinearity, such
as second-order ­ nonlinearity, Raman scattering, and Brillouin
scattering.
Nonlinear Optics Applications of Microring Resonators 237

References
Absil, P. P., Hryniewicz, J. V., Little, B. E., Cho, P. S., Wilson, R. A.,
Joneckis, L. G., Ho, P.-T. 2000. Wavelength conversion in GaAs
micro-ring resonators. Opt. Lett. 25: 554–556.
Agrawal, G. P. 2013. Nonlinear Fiber Optics, 5th Ed. Oxford: Academic
Press.
Almeida, V. R., Barrios, C. A., Panepucci, R. R., Lipson, M. 2004.
All-optical control of light on a silicon chip. Nature 431(28):
1081–1084.
Armaroli, A., Malaguti, S., Bellanca, G., Trillo, S., de Rossi, A.,
Combrié, S. 2011. Oscillatory dynamics in nanocavities with
noninstantaneous Kerr response. Phy. Rev. A 84(5): 053816.
Boyd, R. W. 2008. Nonlinear Optics, 3rd Ed. New York: Academic
Press.
Chen, S., Zhang, L., Fei, Y., Cao, T. 2012. Bistability and self-pulsation
phenomena in silicon microring resonators based on nonlinear
optical effects. Opt. Express 20(7): 7454–7468.
Clemmen, S., Phan H. K., Bogaerts, W., Baets, R. G., Emplit, Ph.,
Massar, S. 2009. Continuous wave photon pair generation in sil-
icon-on-insulator waveguides and ring resonators. Opt. Express
17(19): 16558–16570.
Del’Haye, P., Arcizet, O., Schliesser, A., Holzwarth, R., Kippenberg,
T. J. 2008. Full stabilization of a microresonator-based optical
frequency comb. Phys. Rev. Lett. 101: 053903.
Dimitropoulos, D., Jhaveri, R., Claps, R., Woo, J. C. S., Jalali, B. 2005.
Lifetime of photogenerated carriers in silicon-on-insulator rib
waveguides. Appl. Phys. Lett. 86: 071115.
Dinu, M., Quochi, F., Garcia, H. 2003. Third-order nonlineari-
ties in silicon at telecom wavelengths. Appl. Phys. Lett. 82(18):
2954–2956.
Eggleton, B. J., Moss, D. J., Radic, S. 2008. Optical Fiber Telecommunica-
tions V: Components and Sub-systems (eds. I. P. Kaminow, T. Li,
A. E. Willner), Chapter 20, pp. 759–828. San Diego: Academic
Press.
238 Optical Microring Resonators

Ibrahim, T. A., Amarnath, K., Kuo, L. C., Grover, R., Van, V., Ho,
P.-T. 2004. Photonic logic NOR gate based on two symmetric
microring resonators. Opt. Lett. 29(23): 2779–2781.
Ibrahim, T. A., Cao, W., Kim, Y., Li, J., Goldhar, J., Ho P.-T., Lee, C. H.
2003a. All-optical switching in a laterally coupled microring
resonator by carrier injection. IEEE Photonics Technol. Lett. 15(1):
36–38.
Ibrahim, T. A., Grover, R., Kuo, L.-C., Kanakaraju, S., Calhoun,
L. C., Ho, P.-T. 2003b. All-optical AND/NAND logic gates
using ­semiconductor microresonators. IEEE Photonics Technol.
Lett. 15(10): 1422–1424.
Ikeda, K., Akimoto, O. 1982. Instability leading to periodic and
­chaotic self-pulsations in a bistable optical cavity. Phys. Rev.
Lett. 48(9): 617–620.
Ikeda K., Daido, H. 1980. Optical turbulence: Chaotic behavior
of transmitted light from a ring cavity. Phys. Rev. Lett. 45(9):
709–712.
Ikeda, K., Saperstein, R. E., Alic, N., Fainman, Y. 2008. Thermal and
Kerr nonlinear properties of plasma-deposited silicon nitride/
silicon dioxide. Opt. Express 16(17): 12987–12994.
Islam, M. N., Soccolich, C. E., Slusher, R. E., Levi, A. F. J., Hobson,
W. S., Young, M. G. 1992. Nonlinear spectroscopy near half-gap
in bulk and quantum well GaAs/AlGaAs waveguides. J. App.
Phys. 71(4): 1927–1935.
Johnson, T. J., Borselli, M., Painter, O. 2006. Self-induced ­optical
modulation of the transmission through a high-Q silicon
­
microdisk resonator. Opt. Express 14: 817–831.
Lin, Q., Painter, O. J., Agrawal, G. P. 2007. Nonlinear optical phe-
nomena in silicon waveguides: Modeling and applications.
Opt. Express 15(25): 16604–16644.
Malaguti, S., Bellanca, G., de Rossi, A., Combrie, S., Trillo, S. 2011.
Self-pulsing driven by two-photon absorption in semiconduc-
tor cavities. Phys. Rev. A 83: 051802(R).
Marhic, M. E., Kagi, N., Chiang, T.-K., Kazovsky, L. G. 1996.
Broadband fiber optical parametric amplifiers. Opt. Lett. 21(8):
573–575.
Nonlinear Optics Applications of Microring Resonators 239

Marcuse, D. 1991. Theory of Dielectric Optical Waveguides, 2nd


Ed. Boston: Academic Press.
Ong, J. R., Kumar, R., Aguinaldo, R., Mookherjea, S. 2013. Efficient
CW four-wave mixing in silicon-on-insulator micro-rings
with active carrier removal. IEEE Photonics Technol. Lett. 25(17):
1699–1702.
Pernice, W. H. P., Li, M., Tang, H. X. 2010. Time-domain measure-
ment of optical transport in silicon micro-ring resonators.
Opt. Express 18: 18438–18452.
Priem, G., Dumon, P., Bogaerts, W., Van Thourhout, D., Morthier, G.,
Baets, R. 2005. Optical bistability and pulsating behaviour in
silicon-on-insulator ring resonator structures. Opt. Express 13:
9623–9628.
Razzari, L., Duchesne, D., Ferrera, M., Morandotti, R., Chu, S., Little,
B. E., Moss, D. J. 2010. CMOS-compatible integrated optical
hyper-parametric oscillator. Nat. Photonics 4: 41–45.
Silberberg, Y., Bar-Joseph, I. 1984. Optical instabilities in a nonlinear
Kerr medium. J. Opt. Soc. Am. B 1(4): 662–670.
Soref, R., Bennett, B. 1987. Electrooptical effects in silicon. IEEE
J. Quantum Electron. 23: 123–129.
Turner-Foster, A. C., Foster, M. A., Levy, J. S., Poitras, C. B., Salem,
R., Gaeta, A. L., Lipson, M. 2010. Ultrashort free-carrier life-
time in low-loss silicon nanowaveguides. Opt. Express 18(4):
3582–3591.
Van, V., Ibrahim, T. A., Absil, P. P., Johnson, F. G., Grover, R., Ho, P.-T.
2002a. Optical signal processing using nonlinear semiconduc-
tor microring resonators. IEEE J. Sel. Top. Quantum Electron. 8(3):
705–713.
Van, V., Ibrahim, T. A., Ritter, K., Absil, P. P., Johnson, F. G., Grover,
R., Goldhar, J., Ho, P.-T. 2002b. All-optical nonlinear switching
in GaAs-AlGaAs microring resonators. IEEE Photonics Technol.
Lett. 14(1): 74–77.
Van, V., Little, B. E. 1999. Design and analysis of nonlinear
m icroring resonators for third-order harmonic generation.
­
In Integrated Photonics Research Conference, Santa Barbara, CA,
paper RTuB3.
240 Optical Microring Resonators

Zhang, L., Fei, Y., Cao, T., Cao, Y., Xu, Q., Chen, S. 2013. Multibistability
and self-pulsation in nonlinear high-Q silicon microring
resonators considering thermo-optical effect. Phys. Rev. A 87:
053805.
Zhang, L., Fei, Y., Cao, Y., Lei, X., Chen, S. 2014. Experimental obser-
vations of thermo-optical bistability and self-pulsation in
­silicon microring resonators. J. Opt. Soc. Am. B 31(2): 201–206.
Chapter 5

Active Photonic Applications


of Microring Resonators

In Chapter 4, we showed that the resonant frequencies of a micror-


ing resonator can be modified by the electric field of the lightwave
circulating inside the resonator through the nonlinear optical pro-
cesses of self-phase modulation and cross phase modulation. In
this ­chapter, we consider microring devices in which the resonant
frequency can be tuned by applying an electrical signal such as a
voltage or a current. The ability to electrically tune the microring
resonances allows a variety of important active photonic device
applications to be realized such as tunable filters, switches, and
modulators.
This chapter reviews the principles and technologies for
­achieving tuning, switching, and modulation of a microring res-
onator by modifying its resonant frequencies. Section 5.1 gives
an overview of the common physical mechanisms by which
the resonant frequencies of a microring resonator can be tuned,
namely, the thermo-optic effect, the electro-optic effect, and free
carrier dispersion. The dynamic response of a microring modula-
tor under the small-signal condition will be analyzed in Section
5.2. Section 5.3 extends the analysis to large signals by taking
into account the nonlinear transfer function of the microring
modulator. Important device performance characteristics such as
modulation efficiency, bandwidth, and nonlinearity will also be
discussed.

5.1 Mechanisms for Tuning Microring Resonators


In general, to tune the resonant wavelengths of a microring reso-
nator, we need to change the effective index neff of the microring
waveguide. Suppose the effective index of a section of length L of
the microring is changed by Δneff, for example, by the application of

241
242 Optical Microring Resonators

a voltage V,* we can determine the wavelength shift Δλ experienced


by resonance mode λm as follows. We begin by writing the reso-
nance condition as

(2πR − L)neff (λ ) + Lneff (λ , V ) = mλ m , (5.1)

where we have separated the microring circumference into a


section of length L whose effective index depends on both the
wavelength and the applied voltage, and a section of length
­
(2πR − L) whose effective index is not affected by the voltage.
Taking the ­differential change of Equation 5.1 with respect to both
the ­wavelength and the applied voltage, we get

dneff  ∂n ∂n 
(2πR − L) ∆λ + L  eff ∆λ + eff V  = m∆λ , (5.2)
dλ  ∂λ ∂V 

which can be simplified to give

 dn n  ∂n
2πR  eff − eff  ∆λ + L eff V = 0, (5.3)
 dλ λm  ∂V

where we have made use of the relation m = 2πRneff/λm. Using the


expression for the group index, ng = neff − λmdneff/dλ, we obtain from
Equation 5.3

λm L
∆λ = ∆neff (V ), (5.4)
ng 2πR

where Δneff(V) = (∂neff/∂V)V is the index change due to the applied


voltage alone.
We can vary the effective index of the microring waveguide by
changing the refractive index of the core material, the cladding, or
both. If Δnc and Δncl are the index changes in the core and cladding
material, respectively, the resulting change in the effective index
can be evaluated from the expression

2nc ∆nc Γ c 2ncl ∆ncl Γ cl


∆neff = + , (5.5)
neff neff

* The analysis also holds if the voltage is replaced by any other external means of
changing the effective index, such as temperature, pressure, or electric field in
the case of nonlinear optical waveguides.
Active Photonic Applications of Microring Resonators 243

where nc and ncl are the original indices of the core and cladding,
respectively. The factors Γc and Γcl represent the fraction of the
mode power residing in the core and cladding, respectively, and
are given by

∫ φ(x, y) dx dy ,
2

Si
Γ =
i (5.6)
∫ φ(x, y) dx dy
2

where i = {c, cl}, ϕ(x, y) is the field distribution of the waveguide


mode, and Si is the cross-sectional area of the core or cladding.
The changes in the refractive indices of the core and cladding
materials can be effected by several mechanisms, the most com-
mon being the thermo-optic effect, the electro-optic effect, and free
­carrier dispersion in the waveguide. We briefly discuss each of
these tuning mechanisms below.

5.1.1 The thermo-optic effect


The thermo-optic effect describes the change in the refractive index
of a material with temperature, which can be expressed by the
formula

n(T ) = n(T0 ) + α(T − T0 ), (5.7)

where α = dn/dT is the thermo-optic coefficient. For example, the


thermo-optic coefficients of Si and SiO2 at room temperature are
1.86 × 10−4 K−1 (Cocorullo et al. 1999) and 9 × 10−6 K−1 (Leviton and
Frey 2006), respectively. In general, the thermo-optic coefficients
themselves are also functions of the temperature.
Since both the core and cladding materials of a waveguide
exhibit the thermo-optic effect, the variation in the effective index
of the waveguide with temperature will be the result of the thermo-
optic changes in the indices of both the core and cladding mate-
rials, as given by Equation 5.5. Some materials, such as SU8 and
PMMA (polymethylmethacrylate) polymers, have negative thermo-
optic coefficients. These materials can be used as the cladding to
­compensate for the positive thermo-optic change in the r­ efractive
index of the core material in order to minimize the tempera-
ture ­sensitivity of the device, or even to achieve athermal device
­operation (Raghunathan et al. 2010).
244 Optical Microring Resonators

A simple way to achieve thermo-optic tuning of a microring


resonator (or any integrated optics device) is by varying the temper-
ature of the substrate using a Peltier cooler or heater. This method is
useful for controlling the temperature of the entire chip. For more
localized thermo-optic control of a microring, a micro-heater is typ-
ically fabricated directly above the structure, as shown in Figure 5.1.
Micro-heaters are typically made of highly resistive materials, such
as metals (e.g., Ni/Cr [Sherwood-Droz et al. 2008], Ti [Dong et al.
2010a]). By passing a current through the heater element, resistive
dissipation causes local heating of the waveguide section directly
underneath the heater. The temperature change experienced by
the waveguide can be determined by solving the heat diffusion
equation

∂T
ρC = ∇ ⋅ ( K ∇T ) + q , (5.8)
∂t

over the waveguide cross section in the x–y plane, as shown in


Figure 5.1b. In the above equation, T(x, y) is the temperature distri-
bution, ρ, C, and K are the mass density, specific heat capacity, and
thermal conductivity, respectively, of the materials comprising the
waveguide. The heat source q can be equated to the power density
dissipated from the resistor, q = J2/σ, where J is the current density
and σ is the electric conductivity of the heater. Under steady-state
condition, Equation 5.8 reduces to Poisson’s equation, ∇⋅(K∇T) = −q.
From the temperature distribution, we can calculate the thermo-
optic change in the refractive index n(x, y) over the waveguide cross
section, which then allows us to determine the new effective index

(a) (b) Resistive layer

Micro-heater y
SiO2 cladding

+V –V Si core
x
SiO2 buffer

Si substrate

Bus waveguide Microring

Figure 5.1 Schematic of a micro-heater on a silicon microring resonator:


(a) top view and (b) cross-sectional view.
Active Photonic Applications of Microring Resonators 245

of the waveguide. Typically, the effective index has a roughly l­ inear


dependence on the power and can be expressed by the relation
Δneff = ηP/L, where η is the thermo-optic tuning rate, and P is the
total power dissipated over the length L of the micro-heater. Using
Equation 5.4, we find that the shift in the resonant wavelengths of a
microring resonator of radius R is given by

ηP λ m
∆λ = . (5.9)
2πR ng

The above expression shows that the resonant wavelength shift is


directly proportional to the dissipated power and inversely propor-
tional to the microring radius. Thus, for the same amount of power
dissipation, we obtain a larger wavelength shift in a microring with
a smaller radius. However, the power required to tune a resonant
wavelength across the full FSR of a microring is relatively inde-
pendent of its radius.* Note also that the resistor length L does not
appear in Equation 5.9 since it is the total power dissipated, and not
the power dissipated per unit length, that determines the ­resonant
wavelength shift.
In designing micro-heaters, consideration should be paid to
the optimal placement of the heater above the waveguide so as to
maximize the temperature rise with minimum power consump-
tion, while avoiding excess optical loss due to light absorption in
the conductor. Another design issue is thermal crosstalk, which is
important when several components are placed in close proximity
to each other. An example is the thermo-optic tuning of individ-
ual microrings in a coupled microring device. In such a structure,
heating one microring may also cause residual heating of adjacent
resonators, resulting in unwanted resonant shifts in the latter. Thus,
the heaters must be carefully designed and positioned in order to
­minimize thermal crosstalk.
The time response of the thermo-optic effect is limited by
thermal diffusion, which is a relatively slow process, typically
­
on the scale of micro or milliseconds. For this reason, thermo-
optic tuning is more suitable for applications in which high-speed
­tuning or modulation of the microring resonance is not required.

* Since the FSR of a microring resonator is given by ΔλFSR = λ2/2πngR, the power
required to tune a resonant wavelength over the full FSR is P = (ΔλFSR2πR/η) ×
(ng/λ) = λ/η, which is independent of the microring radius.
246 Optical Microring Resonators

Some application examples include thermo-optic switches in a


WDM switch matrix, stabilization of the resonance wavelength of
a microring resonator, and alignment of microring resonances in
high-order microring filters.
Owing to the large thermo-optic coefficient of silicon, it is
­possible to achieve efficient thermal tuning of silicon microring
­resonators. For example, for a silicon waveguide with core dimen-
sions 250 × 450 nm2 lying on a 1-μm-thick buffer oxide, simula-
tion shows that using a NiCr heater with 2 μm width and 100 nm
thickness located 1 μm above the waveguide, we can achieve a
tuning rate of about η = 4.3 × 10−2 μm/mW in terms of the effective
index change of the TE mode. Figure 5.2a shows the temperature
distribution over the waveguide cross section when the power
­dissipation is 0.1 mW per μm of heater length. For a 10-μm-radius
silicon microring resonator, assuming that the micro-heater cov-
ers the entire microring circumference, the tuning rate is around
20 mW/FSR shift in the microring resonant wavelength. Figure 5.2b
shows the tuning curves (wavelength shift vs. power) of the s­ ilicon
microring for different heater positions above the silicon wave-
guide. We observe that the tuning curves are relatively linear,
with the tuning rate increasing as the heater is placed closer to the
microring waveguide. In general, the tuning rate depends on vari-
ous device parameters such as the dimensions of the waveguide
and the micro-heater, the buffer oxide thickness, the location of the
heater above the waveguide, and the polarization of the input light.

(b)
(a) 16
42.143
5000 14
Wavelength shift (nm)

4000 Heater 40
3000 12
35
2000 30 10
y (nm)

1000 25
0 8
–1000 20
–2000 15 6
Si waveguide
–3000 10
5 4
–4000 h = 750 nm
0 2 h = 1000 nm
–5000 0 5000 0 h = 1250 nm
0
x (nm) 0 5 10 15 20 25 30
Power (mW)

Figure 5.2 (a) Temperature distribution (degree celsius) over a ­silicon


waveguide cross section for a 0.1 mW of power dissipation per μm
heater length and (b) resonant wavelength shift of a 10-μm-radius silicon
­microring as a function of the dissipated power for heater position (h) of
750 nm, 1000 nm, and 1250 nm above the silicon waveguide.
Active Photonic Applications of Microring Resonators 247

Experimentally, reported tuning rates for silicon microring resona-


tors are in the range of 10–50 mW/FSR (Sherwood-Droz et al. 2008,
Dong et al. 2010a,b).

5.1.2 The electro-optic effect


The electro-optic effect refers to the change in the refractive index
of a material caused by an applied electric field. In contrast to the
nonlinear optic effects discussed in Chapter 4, the electric field here
is assumed to be static or slowly varying compared to the relax-
ation time of the material. The formalism for describing the electro-
optic effect is based on the optical indicatrix (or index ellipsoid),
which maps the index experienced by a wave propagating in a
given ­direction in a crystal. In the absence of an applied field, the
­equation for the index ellipsoid is

x2 y 2 z2
+ + = 1, (5.10)
nx2 ny2 nz2

where nx, ny, and nz are the indices seen by a wave polarized along
the principal optical axes (x, y, z) of the crystal. The refractive index
seen by a wave with the electric field polarized in a given direction
is found from the intersection of the ellipsoid with the line drawn
from the origin and in the direction parallel to the polarization
vector.
An applied electric field has the effect of rotating the index
ellipsoid, so that the equation for the new ellipse assumes the more
general form (following the convention in Yariv (1991))

 1 2  1 2  1 2  1  1
 2  x +  2  y +  2  z + 2  2  yz + 2  2  zx
n 1 n 2 n 3 n 4 n 5
 1
+2  2  xy = 1. (5.11)
 n 6

The coefficients (1/n2)i are optical constants describing the new


index ellipsoid. They can be related to the coefficients (1/n2 )(i 0 ) of
the ellipsoid in the absence of the applied field by
(0)
 1  1  1
 2  =  2  + ∆  2  , (5.12)
n i n i n i
248 Optical Microring Resonators

where Δ(1/n2)i represents the change in the coefficients due to the


applied field. In the absence of the applied field, Equation 5.11
reduces to Equation 5.10, so we must have
(0) (0) (0)
 1 1  1 1  1 1
 2  = 2 ;  2  = 2 ;  2  = 2 ,
n 1 nx n 2 ny n 3 nz
(0) (0) (0)
 1  1  1
 2  =  2  =  2  = 0.
n 4 n 5 n 6

For the linear electro-optic effect (also known as the Pockels effect),
the changes in the optical constants are directly proportional to the
applied field and are given by
3
 1
∆ 2  =
 n i ∑r
j =1
i, j Ej , (5.13)

where ri,j are the linear electro-optic coefficients and the subscript
j = {1, 2, 3} corresponds to field directions {x, y, z}. Equation 5.13 can
be written in tensor form as

 ∆(1/n2 )1   r11 r12 r13 


 2  
 ∆(1/n )2   r21 r22 r23 
  E1 
 ∆(1/n2 )3   r31 r32 r33   
 =  E2 = r  E, (5.14)
2
 ∆(1/n )4   r41 r42 r43   
E3 
 ∆(1/n2 )5   r51 r52 r53   
 2
  
 ∆(1/n )6   r61 r62 r63 

where r is the linear electro-optic tensor. The linear electro-


optic effect occurs only in crystals with no inversion sym-
metry, such as GaAs, AlGaAs, GaP, and LiNbO3. On the other
hand, centrosymmetric crystals such as Si exhibit the quadratic
electro-­optic effect. Even in crystals lacking inversion symmetry,
rotational symmetry considerations further lead to a reduction
in the number of unique nonzero elements in the electro-optic
tensor. For example, the electro-optic tensor of GaAs has only
three ­nonzero ­elements, r41 = r52 = r63 = 1.43 pm/V, at the 1.15 μm
wavelength (Yariv 1991). The equation for the index ellipsoid
of a GaAs ­crystal in the ­presence of an applied electric field
E = Ex xˆ + Ey yˆ + Ez zˆ is
Active Photonic Applications of Microring Resonators 249

1 2 1 2 1 2
x + 2 y + 2 z + 2r41Ex yz + 2r41Ey zx + 2r41Ez xy = 1, (5.15)
n02 n0 n0

where n0 = 3.4 is the refractive index of GaAs when there is no


applied field.
The refractive index of a crystal in the presence of an applied
field is found from the intersection of the new ellipsoid with the
line drawn from the origin and in the direction parallel to the
wave’s polarization. For example, suppose a GaAs crystal is subject
to an electric field in the x direction, (E = Ex xˆ ), the equation for the
new index ellipsoid is

 1 2  1 2  1 2
 n2  x +  n2  y +  n2  z + 2r41Ex yz = 1. (5.16)
0 0 0

Making the coordinate transformations


X = x,
Y = y cos( 45°) + z sin( 45°) = ( y + z)/ 2 ,
Z = − y sin( 45°) + z cos( 45°) = ( z − y )/ 2 ,

we can write Equation 5.16 as

X2  1   1 
+  2 − r41Ex  Y 2 +  2 + r41Ex  Z 2 = 1. (5.17)
n0  n0
2
  n0 

By equating

1 1
2
= 2 − r41Ex ,
nY n0
1 1
2
= 2 + r41Ex ,
nZ n0

we find that the refractive indices seen by a wave polarized in the


Y and Z directions, respectively, are given by

1 3
nY ≈ n0 + n0 r41Ex , (5.18)
2

1
nZ ≈ n0 − n03 r41Ex . (5.19)
2
250 Optical Microring Resonators

Note that the index seen by a wave polarized in the X direction


is nX = n0, indicating that the wave is unaffected by the applied
field. This example demonstrates that the electro-optic effect is
highly anisotropic and can even vanish for certain directions of
polarization.
In electro-optic modulator applications, an important param-
eter commonly used to quantify the strength of the nonlinear
effect is Vπ, which is the voltage required to achieve a π phase shift
in the lightwave. For example, for a GaAs waveguide of length L
placed between two electrodes separated by distance d, the voltage
required to induce a π phase shift in the lightwave is Vπ = λd/n03 r41L.
Since the electro-optic effect has its origin in the response
of bound electrons to the applied electric field, it is character-
ized by a very fast response time, typically in the order of 10−14 s
(Agullo-Lopez et al. 1994). For this reason, the electro-optic effect
is commonly employed in high-speed modulator applications.
The switching speed of an electro-optic modulator is limited not by
the material response time, but rather by the parasitic RC time con-
stant of the electrical connections, as well as the velocity mismatch
between the electrical and optical signals.
Materials exhibiting the electro-optic effect are not restricted to
only crystalline solids. Certain classes of polymers are also known
to possess very large electro-optic coefficients. For example, a
­phenyltetraene-bridged chromophore (labeled CLD) in an amorphous
polycarbonate (APC) host material has an electro-optic ­coefficient of
55 pm/V at the 1.55 μm wavelength (Oh et al. 2000), which is almost
40 times larger than that in GaAs. This polymer material has also
been used to fabricate efficient microring modulators operating in
the telecommunication wavelength range (Rabiei et al. 2002).

5.1.3 Free carrier dispersion


The effective index of a semiconductor waveguide can also be mod-
ulated by varying the free carrier concentration in the ­waveguide.
The change in the refractive index and optical absorption of a
­semiconductor material due to the presence of Ne electron concentra­
tion and Nh hole concentration are given by Equations 4.16 and 4.17

∆n = σ (re ) N e + σ (rh ) N h , (5.20)

∆α = σ (ae ) N e + σ (ah ) N h , (5.21)


Active Photonic Applications of Microring Resonators 251

where σ (re , h ) and σ (ae , h ) are the refraction volume and absorption cross
section, respectively, due to electron (e) and hole (h) concentrations.
The refraction volume and absorption cross section may be evalu-
ated by treating the electrons and holes as plasmas embedded in a
background dielectric of index n0. The relative permittivity of the
semiconductor in the presence of excess electrons or holes can be
written as

ε r + ∆ε r = (n0 + ∆n − j∆α )2 ≈ n02 + 2n0 (∆n − j∆α ), (5.22)

where ε r = n02. From the Drude model for a plasma, the change in
the relative permittivity Δεr due to electron or hole concentration
N(e,h) is given by

N( e , h ) q 2 1
∆ε r ≈ − ∗
, (5.23)
ε 0 m( e , h ) ω(ω − jγ )


where m(e,h) is the effective mass of an electron or hole in the
­semiconductor, γ is the free carrier relaxation rate, and q is the
­elementary charge. Substituting Equation 5.23 into Equation 5.22
and using the relations in Equations 5.20 and 5.21, we get

(e,h ) q2  1 
σ (ω) ≈ − , (5.24)
2n0 ε 0 m( e , h )  ω + γ 2 

r ∗ 2

q2  γ /ω 
σ (ae , h ) (ω) ≈ . (5.25)
2n0 ε 0 m( e , h )  ω 2 + γ 2 

For silicon around the 1.55 μm wavelength, it is more accurate to


use the empirical values for σ (re , h ) and σ (ae , h ) given in Equations 4.20
and 4.21.
Free carriers are typically induced in a semiconductor wave-
guide by the application of a voltage across a semiconductor
­junction. A variety of semiconductor junctions have been employed
to achieve modulation of the free carrier density, including Schottky
junction (Campbell et al. 1975), pn junction (Dong et al. 2009), pin
junction (Xu et al. 2005, Li et al. 2007), metal–oxide–­semiconductor
capacitor (MOS CAP) (Liu et al. 2004), and silicon–insulator–­silicon
capacitor (SISCAP) (Milivojevic et al. 2013). Figure 5.3 shows the
schematics of silicon waveguide structures embedded with a
252 Optical Microring Resonators
(a) (b)
Contact Contact Contact Contact
SiO2 SiO2
250 nm 250 nm
n+ Si p+ Intrinsic p+
n p n+
Si
SiO2 80 nm SiO2 50 nm

Si substrate Si substrate

(c)
Contact Contact

SiO2 p-Si

Gate oxide n-Si

SiO2
Si substrate

Figure 5.3 Schematic of (a) a horizontal pn junction silicon waveguide


(After Dong, P., et al., 2009, Opt. Express, 17, (25): 22484–22490), (b) a pin
junction silicon waveguide (After Xu, Q., et al., 2005, Nature, 435, (19):
325–327), and (c) an MOS capacitor silicon rib waveguide (After Liu, A.,
et al., 2004, Nature, 427, (12): 615–618).

pn junction, a pin junction, and an MOS capacitor. Typically, pin


junctions are operated in the forward bias mode (carrier injection),
while pn junctions can be operated in either the forward bias or
reverse bias mode (depletion mode). Likewise, MOS capacitors can
be ­operated in either the carrier depletion mode or carrier accu-
mulation mode. In general, the forward bias mode provides larger
changes in the free carrier densities and hence larger changes in
the refractive index than the reverse bias mode. However, it is also
inherently slower since a larger amount of charge must be moved
in and out of the junction. For this reason, high-speed modula-
tors ­typically employ depletion-mode pn junctions or depletion-
mode MOS capacitors. The ­switching or modulation speed of these
­junctions is determined by the recombination time and transit time
of the free carriers, as well as the RC time constant of the electrical
contacts.
The relationship between the induced free carrier concentra-
tions and applied voltage in a semiconductor junction waveguide is
determined by solving Poisson’s equation of electrostatics and the
charge transport equation in the ­junction (Sze 1981). Commercial
Active Photonic Applications of Microring Resonators 253
(a) (b)
Microring Metal contacts

Pout(t)

vm(t)

Pin Pout(t)
ω0 ω
δωm(t)

Figure 5.4 (a) Schematic of an all-pass microring modulator with


­constant input light power Pin and modulating voltage vm(t) applied across
the microring waveguide and (b) plot of the microring resonance spec-
trum showing modulation in the transmitted power Pout(t) due to varia-
tion in the resonant frequency δωm(t).

software such as Atlas, Crosslight, Lumerical, and Comsol, are


also available which can be used to obtain accurate simulations of
the charge–voltage characteristics and the time-domain response
of charge transport in practical device structures. Electrical cir-
cuit models of the junctions can also be constructed to assist in
the design and optimization of switches and modulators based on
these semiconductor junction waveguides.

5.2 Dynamic Response of a Microring Modulator


In this section, we develop a set of equations based on the energy
coupling formalism for analyzing the dynamic response of an
APMR due to a change in the resonant frequency (Pile and Taylor
2014). These equations are general in that they are not specific to the
underlying mechanism used to effect the resonant frequency shift,
and can be used to model the dynamic response of a microring in
switching and modulation applications. Since these applications
typically involve small shifts in the resonant frequency compared
to the FSR of the resonator, the energy coupling formalism will
­provide sufficient accuracy for our analysis.*
The principle of operation of an APMR switch or intensity
modulator is illustrated in Figure 5.4. An input monochromatic

* An analysis of the dynamic response of a microring modulator based on the


power coupling formalism can be found in Sacher and Poon (2008).
254 Optical Microring Resonators

lightwave is tuned to a frequency on the shoulder of the resonance


spectrum of the microring resonator. Suppose a voltage signal vm(t)
is applied across the microring waveguide which induces a change
in its ­effective index. This causes a shift δωm in the resonance
­spectrum of the microring resulting in a corresponding change in
the transmitted light power, Pout(t), of the microring resonator.
In Section 5.2.1, we will formulate a set of equations based on
energy coupling which govern the time-domain response of an
APMR modulator. We will then solve these equations in the small-
signal (SS) approximation in Section 5.2.2, where the modulator
response is ­linearized around a bias point. Important characteristics
of the modulator such as modulation efficiency, bandwidth, and step
response will be discussed. Extension of the analysis to large-signal
(LS) condition, which takes into account the nonlinear characteristic
of the microring transfer function, will be treated in Section 5.3.

5.2.1 Dynamic energy-coupling model


of a microring modulator
We consider an APMR with resonant frequency ω0, energy cou-
pling coefficient μ, and intrinsic decay rate γ0 due to loss. The
­intrinsic decay rate can be computed from the microring waveguide
loss α as γ0 = αvg/2, where vg is the group velocity. For an optical sig-
nal sin (t) applied to the input port of the APMR, the equation for the
energy-normalized wave amplitude a(t) in the microring is
da
= ( jω 0 − γ r )a(t) − jµsin (t), (5.26)
dt
where γr = γ0 + γe is the total decay rate and γe = μ2/2 is the extrinsic
decay rate due to coupling to the bus waveguide. In the following
analysis, we assume that the input light is a monochromatic wave
with frequency ω and power Pin so that sin (t) = Pin e jωt. Supposing
that the resonant frequency of the microring experiences a time-
dependent change given by δωm(t), the instantaneous microring
resonant frequency is ω0 + δωm. Writing a(t) = a(t)e jωt, we obtain
from Equation 5.26
da
= −( j∆ω + γ r )a + jδω m a − jµ Pin , (5.27)
dt
where Δω = ω − ω0 is the frequency detune of the input light signal
from the microring resonance.
Active Photonic Applications of Microring Resonators 255

The change in the microring resonant frequency is related to


the change in the effective index δnr of the microring waveguide by
δωm(t) = −ω0δnr(t)/ng, where ng is the group index. We will assume
that the effective index can be changed through the application of a
voltage vm to the modulator, and that the relationship between the
index change and the applied voltage is linear, δnr = ηvm, where η
is the refractive index change per volt. We will further assume that
the response time of the effective index to the applied voltage is
characterized by a time constant τm, which we call the index mod-
ulation time constant. This time constant depends on the specific
mechanism used to induce the index change, the physical structure
and dimensions of the device, and the parasitic capacitances and
resistances associated with the electrical contacts. With the above
assumptions, we can express the time dependence of the resonant
frequency modulation on the applied voltage as

dδω m δω  ηω 0 
= − m − vm (t). (5.28)
dt τ m  ng τ m 

If the response of the effective index to the applied voltage is instan-


taneous, then Equation 5.28 simply gives δωm(t) = −(ηω0/ng)vm(t). The
solution of Equations 5.27 and 5.28 gives the dynamic response of
the wave amplitude a(t) in the microring. The transmitted light sig-
nal at the output of the APMR can then be determined from the
relation sout (t) = Pin − jµa(t).
The above equations can also be applied to a microring modula-
tor or switch in the add-drop configuration. For an ADMR with input
and output energy coupling coefficients μ1 and μ2, respectively, the
total decay rate in Equation 5.27 is given by γ r = γ 0 + µ12 /2 + µ 22 /2 and
the input term is modified to read − jµ1 Pin . The transmitted light
signals at the drop port and through port of the microring are given
by the relations sd(t) = −jμ2a(t) and st (t) = Pin − jµ1 a(t), respectively.
In the next section, we will solve Equations 5.27 and 5.28 for
an APMR modulator under the SS approximation to obtain the
relationship between the output modulated light power and the
applied voltage.

5.2.2 Small-signal analysis


In the absence of an applied modulating signal (δωm = 0), the wave
amplitude in the microring is constant, a(t) = a0, with the solution
given by
256 Optical Microring Resonators

− jµ Pin − jµ Pin
a0 = = , (5.29)
j∆ω + γ r Γ

where Γ = jΔω + γr. Suppose a small modulating signal vm(t) is


applied to the microring causing a small resonant frequency shift
δωm(t). As a result, the wave amplitude in the microring deviates
from the steady-state value by a small amount δam(t) so that we can
write a(t) = a0 + δam(t). Making this substitution in Equation 5.27 and
keeping only terms linear in δωm(t) and δam(t), we get

dδam
= −Γδam + ja0 δω m . (5.30)
dt

Taking the Fourier transform of the above equation and solving for
δam, we obtain

ja0
δam ( jΩ) = δω m ( jΩ), (5.31)
jΩ + Γ

where Ω represents the electrical (or RF) modulation frequency,


and δam(jΩ) and δωm(jΩ) denote the Fourier transform of the wave
amplitude in the microring and the resonant frequency modula-
tion, respectively. The resonant frequency spectrum δωm(jΩ) can be
obtained from the Fourier transform of Equation 5.28 as

 ηω 0  Vm ( jΩ)
δω m ( jΩ) = −   , (5.32)
 ng τ m  jΩ + 1/τ m

where Vm(jΩ) denotes the Fourier transform of the modulating


­ oltage signal. Substituting Equation 5.32 into Equation 5.31, we get
v

 ηω 0  ja0Vm ( jΩ)
δam ( jΩ) = −   . (5.33)
 ng τ m  ( jΩ + Γ )( jΩ + 1/τ m )

Since the transmitted light signal at the output of the APMR


is given by sout (t) = sin − jµa(t), we can determine the output modu-
lated light power from
2 2
Pout (t) = sout (t) = Pin − jµa(t) . (5.34)
Active Photonic Applications of Microring Resonators 257

Substituting a(t) = a0 + δam(t) into the above equation, we get


2 2
Pout (t) = Pin − jµ( a0 + δam ) = sout − jµδam , (5.35)

where sout is the steady-state (or DC) transmitted light amplitude in


the absence of modulation

sout = Pin − jµa0 = (1 − µ 2 /Γ ) Pin . (5.36)

We can expand Equation 5.35 to read


2 ∗ ∗ 2
Pout (t) = sout + jµδam sout − jµδam sout DC
+ µ 2 δam = Pout AC
+ Pout (t),
(5.37)
DC
where Pout is the steady-state (or DC) transmitted light power in the
absence of modulation

2 2
DC
Pout = sout = 1 − µ 2 /Γ Pin , (5.38)

AC
and Pout (t) is the AC component of the transmitted light power due
to modulation
∗ ∗ 2
AC
Pout (t) = jµ(δam sout − δam sout ) + µ 2 δam . (5.39)

Upon substituting the expression for sout in Equation 5.36 into


Equation 5.39 and neglecting the small contribution μ2|δam|2, we get

AC
 µ2  ∗  µ2  
Pout (t) = jµ Pin  1 −  δam (t) −  1 − ∗  δam (t) . (5.40)
 Γ  Γ  

Taking the Fourier transform of the above equation, we have

AC
 µ2  ∗  µ2  
Pout ( jΩ) = jµ Pin  1 −  δam (− jΩ) −  1 − ∗  δam ( jΩ) . (5.41)
 Γ  Γ  

By substituting the expression for δam(jΩ) in Equation 5.33 into the


above equation, we obtain the frequency response of the m­ odulator
due to an applied RF signal Vm(jΩ).
258 Optical Microring Resonators

5.2.2.1 Transfer function of an APMR modulator


The transfer function of an APMR modulator can be obtained by
applying an impulse voltage signal vm(t) = δ(t) to the device and
determining its frequency response. Since the Fourier transform of
the impulse signal is Vm(jΩ) = 1, we get from Equation 5.33

 ηω 0  µ Pin /Γ
δam ( jΩ) = −   . (5.42)
 ng τ m  ( jΩ + Γ )( jΩ + 1/τ m )

Substituting the above result into Equation 5.41, we obtain the trans-
fer function of the modulator as

AC K 0 ( jΩ + 2γ 0 )
H m ( jΩ) = Pout ( jΩ) = , (5.43)
( jΩ + Γ )( jΩ + Γ ∗ )( jΩ + 1/τ m )

2∆ωµ 2  ηω 0 
K0 = − Pin .
∆ω 2 + γ r2  ng τ m 

Equation 5.43 indicates that the modulator has a third-order


response characterized by three poles and one zero. Two of the
poles are related to the microring resonance through the complex
frequency parameter Γ, and the third pole arises from the time
response of the waveguide index to the applied voltage, which is
characterized by the index modulation time constant τm.

5.2.2.2 Modulation efficiency


The modulation efficiency of an APMR modulator can be defined as
the modulation gain at DC, G = Hm(0), which is the slope of the trans-
mitted light power versus applied voltage curve at DC. Evaluating
Equation 5.43 at Ω = 0, we get

2γ 0 K 0 4∆ωγ 0 µ 2  ηω 0 
G = H m (0 ) = = − Pin . (5.44)
|Γ|2 (1/τ m ) (∆ω 2 + γ r2 )2  ng 

Note that the DC gain is independent of the index modulation time


constant τm. It is convenient to express Equation 5.44 in terms of
normalized parameters as follows:

 ηP  δ γ
G = −16Q0  in  2 2 3
, (5.45)
 ng  (1 + δ ) (1 + γ )
Active Photonic Applications of Microring Resonators 259

where δ = Δω/γr is the normalized frequency detune, γ = γe/γ0 is


the ratio of the extrinsic decay rate to the intrinsic decay rate, and
Q0 = ω0/2γ0 is the intrinsic Q factor of the resonator. The normal-
ized frequency detune δ gives the ratio of the frequency detune to
the half width at half max (HWHM) bandwidth of the ­microring,
δ = Δω/Δω3dB (since the HWHM bandwidth Δω3dB = γr). The param-
eter γ can also be regarded as the normalized coupling because it
is equal to the energy coupling ­normalized by the intrinsic loss,
γ = μ2/αvg.
From Equation 5.45, we find that for a given intrinsic Q 0 factor
of the microring, the maximum gain, Gmax, occurs at δ = 1/ 3 and
γ = 1/2 and is equal to

4Q0  ηPin 
Gmax = − . (5.46)
3 3  ng 

If we normalize the gain G in Equation 5.45 by Gmax, we obtain the


expression

G 12 3δ γ
= , (5.47)
Gmax (1 + δ 2 )2 (1 + γ )3

which is independent of the microring loss. In general, for a given


value of γ, we obtain the highest DC gain if the normalized fre-
quency detune is set to be δ = 1/ 3 . On the other hand, for a fixed
frequency detune δ, we obtain the maximum DC gain when the
normalized coupling parameter γ is equal to 1/2.
Figure 5.5a and b plot the normalized DC gain (G/Gmax) as a func-
tion of the normalized frequency detune δ and the normalized cou-
pling γ. In Figure 5.5a, we see that the DC gain initially increases with
the frequency detune, reaching a peak value at δ = 1/ 3 ≈ 0.58 before
decreasing again with further increase in the frequency detune.
Note that the choice of δ also determines the optical insertion loss
of the APMR modulator. At δ = 1, which corresponds to an insertion
loss of 3 dB, the DC gain is about 1.1 dB below the peak gain value.
Figure 5.5b shows a similar dependence of the DC gain on the
normalized coupling parameter, with the peak gain value reached
at γ = 1/2. At critical coupling (γ = 1), the DC gain is about 0.7 dB
below the peak gain value. The plot also indicates that in general, to
obtain high modulation efficiency, the microring resonator should
be operated in the undercoupling regime (γ < 1).
260 Optical Microring Resonators
(a) (b)
0 0
Normalized DC gain, G/Gmax (dB)

Normalized DC gain, G/Gmax (dB)


–5
–5

–10

–10
–15 γ = 0.1
γ = 0.5 δ = 0.58
γ=1 δ=1
γ=5 δ = 1.5
–20 –15
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0.1 1 10
Normalized frequency detune, δ Normalized coupling, γ

Figure 5.5 (a) Plot of the normalized DC gain (G/Gmax) versus the nor-
malized frequency detune δ for various values of the coupling parameter
γ and (b) plot of the normalized DC gain versus the normalized coupling
γ for several values of the frequency detune δ.

5.2.2.3 Electrical bandwidth
To determine the electrical (or RF) bandwidth of an APMR modu-
lator, we consider two separate cases: the case where the modula-
tor response is dominated by the index modulation time constant
τm, and the case where it is dominated by the microring resonance.
In the first case, the transfer function of the modulator can be
­simplified by making the approximation |Γ| ≫ Ω in Equation 5.43.
This results in the expression

G/τ m
H m ( jΩ) = , (5.48)
jΩ + 1/τ m

where the DC gain G is given by Equation 5.44. Equation 5.48 indi-


cates that the modulator has a first-order frequency response, with
the 3 dB bandwidth of the magnitude response, |Hm(jΩ)|, given by
Ω 3dB = 3 /τ m.
For the case where the response of the modulator is dominated
by the microring resonance, we make the approximation 1/τm ≫ Ω
in Equation 5.43 to obtain the following transfer function for the
modulator:

2
GΓ ( jΩ + 2γ 0 )
H m ( jΩ) = . (5.49)
2γ 0 ( jΩ + Γ )( jΩ + Γ ∗ )
Active Photonic Applications of Microring Resonators 261

The above transfer function can also be expressed in terms of the


normalized frequency detune and normalized coupling as
2 
 ) = G(1 + γ )(1 + δ )[ jΩ + 2/(1 + γ )] ,
H m ( jΩ (5.50)
 + (1 + jδ )][ jΩ
2[ jΩ  + (1 − jδ )]

where Ω = Ω/γ r = Ω/∆ω 3dB is the normalized RF frequency. The


normalized 3 dB RF bandwidth, Ω  3dB = Ω 3dB /∆ω 3dB , of H m ( jΩ
 ) can
be computed from

 3dB = b + b 2 + 3(1 + δ 2 )2 ,
Ω (5.51)

1
b = (1 + γ )2 (1 + δ 2 )2 − (1 − δ 2 ).
2

When the modulator is operated at maximum gain Gmax (γ = 1/2


and δ = 1/ 3 ), the normalized RF bandwidth is equal to Ω  3dB = 2.
The plots in Figure 5.6a and b show the frequency response of
the normalized gain |H m ( jΩ  )/G| of an APMR modulator for the
case where the index modulation time constant τm can be neglected.
Figure 5.6a plots the frequency responses for a fixed coupling param-
eter γ = 1/2 and various values of the normalized frequency detune,
while Figure 5.6b shows the responses for a fixed frequency detune
δ = 1/ 3 ≈ 0.58 and various values of the coupling parameter. We
observe that the electrical bandwidth of the modulator in general
increases with the frequency detune and the coupling parameter.
The variations of the electrical bandwidth with respect to δ and γ
are further illustrated in Figure 5.6c and d. The increase in the elec-
trical bandwidth with frequency detune can be explained by the
fact that the charging time of the microring resonator decreases as
we move farther away from the microring ­resonance. The micror-
ing charging time can also be decreased by increasing the external
coupling, which lowers the Q factor of the resonator, with the result
that the electrical bandwidth increases with larger γ.
It is also of interest to examine the trade-off between the modu-
lation efficiency and the electrical bandwidth of an APMR modula-
tor. Figure 5.7a plots the relationship between the normalized DC
gain (G/Gmax) and the normalized RF bandwidth (Ω  3dB) at several
fixed values of δ by parameterizing γ. Figure 5.7b plots the same
relationship by parameterizing δ at several fixed values of γ. In both
262 Optical Microring Resonators
(a) (b)
5 5
Normalized RF gain, |Hm/G| (dB)

Normalized RF gain, |Hm/G| (dB)


0 0

–5 –5

–10 –10
γ = 0.1; δ = 0.58
–15 γ = 0.5; δ = 0.58 –15 γ = 0.5; δ = 0.58
γ = 0.5; δ = 1 γ = 1; δ = 0.58
γ = 0.5; δ = 1.5 γ = 5; δ = 0.58
–20 –20
10–3 10–2 10–1 100
101 102 10–3 10–2 10–1 100 101 102
~ ~
Normalized RF frequency, Ω Normalized RF frequency, Ω

(c) (d)
20 25
Normalized RF bandwidth, Ω3dB

Normalized RF bandwidth, Ω3dB


γ = 0.1 δ = 0.58
γ = 0.5 δ=1
~

~
γ=1 20 δ = 1.5
15 γ=5

15
10
10

5
5

0 0
0 0.5 1 1.5 0.1 1 10
Normalized frequency detune, δ Normalized coupling, δ

Figure 5.6 Frequency response of the normalized gain |H m ( jΩ )/G| of an


APMR modulator for (a) fixed coupling parameter γ = 1/2 and various
values of δ, (b) fixed ­frequency detune δ = 1/ 3 and various values of γ,
(c) plot of the ­normalized RF bandwidth versus frequency detune δ for
various ­values of γ, and (d) plot of the normalized RF bandwidth versus
coupling ­parameter γ for various values of δ.

plots, we see that the maximum DC gain is achieved for a normal-


ized RF bandwidth of 2. Further increase in the RF bandwidth
beyond this value leads to a corresponding decrease in the DC gain.
The bandwidth limitation of a microring modulator due to reso-
nance (or the finite cavity lifetime) can be overcome by modulat-
ing the coupling coefficient rather than the round-trip phase of the
APMR (Sacher and Poon 2008, Sacher et al. 2013). The transmitted
light of the APMR is thus modulated by varying the extrinsic Q fac-
tor of the resonator. In this case, the RF gain of the modulator does
not roll off at high frequencies and so, the electrical bandwidth is
not limited by the l­ inewidth of the microring resonance. The trade-
offs of this modulation scheme, however, are the smaller modula-
tion efficiency and larger device size compared to phase-modulated
Active Photonic Applications of Microring Resonators 263
(a) (b)
Normalized DC gain, G/Gmax (dB) 0 0

Normalized DC gain, G/Gmax (dB)


δ = 0.58 γ = 0.2
δ=1 γ = 0.5
–2 δ = 1.5 –2 γ=1
γ=2

–4 –4

–6 –6

–8 –8

–10 –10
0 2 4 6 8 10 12 14 0 2 4 6 8 10
~ ~
Normalized RF bandwidth, Ω3dB Normalized RF bandwidth, Ω3dB

Figure 5.7 Plots of the normalized DC gain (G/Gmax) versus the normal-
ized RF bandwidth (Ω  3dB ) of an APMR modulator by (a) varying γ while
keeping δ fixed and (b) by varying δ while keeping γ fixed.
APMRs (Pile and Taylor 2014). The electrical bandwidth of a cou-
pling-modulated APMR is u
­ ltimately limited by the index modula-
tion time constant τm.

5.2.2.4 Small-signal step response


The SS response of an APMR modulator to a step change in the
­voltage can be obtained by solving Equation 5.30 in the time domain.
Suppose the applied voltage undergoes a step change from 0 to a
constant value V0 at time t = 0. Assuming that the effective index of
the microring waveguide changes instantaneously with the applied
voltage (i.e., τm = 0), the resonant frequency of the microring also
experiences an instantaneous step change equal to Δωm = −(ηω0/ng)V0.
The solution of Equation 5.30 in this case is given by

ja0 ∆ω m
δam (t) = (1 − e − Γt ). (5.52)
Γ

The AC transmitted power of the modulator is found using Equation


5.40*
AC
Pout {
(t) = V0G 1 − [cos(∆ωt) − B sin(∆ωt)] e − γ rt , } (5.53)

* Note that Equation (5.40) neglects the nonlinear term μ2|δam|2, which is propor-
2
tional to ∆ω m (or V02). Since this term is always positive regardless of the sign of
V0, inclusion of this term will result in unequal positive and negative swings
AC
of the transmitted power Pout (t). We omit this term in the SS analysis but will
include it in the LS analysis in Section 5.3.1.
264 Optical Microring Resonators

∆ω 2 + γ e2 − γ 02
B= ,
∆ω/2γ 0

where G is the DC gain in Equation 5.44. The above result shows that
the transmitted power has a second-order time response character-
ized by oscillations at frequency Δω and decay rate γr. Following
the conventional description of a second-order linear system, we
can characterize the SS response of the modulator as overdamped
if Δω/γr = δ < 1, underdamped if Δω/γr > 1, and critically damped
if Δω/γr = 1. Equation 5.53 also indicates that the swing in the
AC
steady-state (or DC) transmitted power, ∆Pout = Pout (t → ∞) = V0G, is
­linearly proportional to the applied voltage V0 (or to the resonant
frequency shift Δωm). The modulation index of the APMR modula-
DC DC
tor can be determined from ηm = V0G/Pout , where Pout is given by
Equation 5.38.

5.3 Large-Signal Response of a Microring Modulator


Since the spectral response of an APMR near a resonance is highly
nonlinear, the transfer function of the microring modulator is
also highly nonlinear. For large modulating signals, this non-
linearity manifests itself through effects such as saturation and
harmonic distortion. In this section, we will investigate the LS
behaviors of an APMR modulator for the cases of input step volt-
age and sinusoidal excitations. In both cases, we will assume that
the effective index of the microring responds instantaneously to
the applied voltage so that we can neglect the index m ­ odulation
time constant τm.

5.3.1 Large-signal step response


To obtain the LS response of an APMR modulator to an applied
step voltage, we solve Equation 5.27 directly in the time domain.
Similar to the SS analysis in the previous section, we assume that
the input voltage undergoes a step change from 0 to V0 at time t = 0,
which causes a corresponding shift in the resonant frequency of
the microring given by Δωm = −(ηω0/ng)V0. The general solution of
Equation 5.27 for the wave amplitude in the microring comprises of
a forced response term and a natural response term, a(t) = af + an(t).
The forced response af is obtained when the modulator has reached
Active Photonic Applications of Microring Resonators 265

a new steady state due to the resonant frequency shift Δωm. Setting
daf/dt = 0 in Equation 5.27, we get

−Γaf + j∆ω m af − jµ Pin = 0,

which can be solved for af to give

− jµ Pin
af = . (5.54)
Γ − j∆ω m

The natural response an(t) is the solution to the homogeneous


equation

dan
= −Γan ,
dt

which yields

an (t) = A0 e − Γt . (5.55)

To determine the constant A0, we note that at t = 0, just before the


step voltage is applied, the wave amplitude in the microring is given
by a(0) = a0 = − jµ Pin /Γ. Using this initial condition to find A0, we
obtain the following solution for the total response a(t):

jµ Pin  j∆ω m − Γt 
a(t) = e − 1 . (5.56)
Γ − j∆ω m  Γ 

To determine the transmitted light power at the modula-


tor ­ output, we write the solution for the wave amplitude as
a(t) = a0 + δam(t), where the LS modulation δam(t) is given by

δam (t) = ∆am (1 − e − Γt ), (5.57)

∆ω m µ Pin
∆am = .
Γ(Γ − j∆ω m )

By substituting δam(t) into Equation 5.39, we obtain the LS mod-


AC
ulation Pout (t) of the transmitted light power, which can be
expressed as
266 Optical Microring Resonators

AC
Pout { }
(t) = V0GLS 1 + [C1 cos(∆ωt) + C2 sin(∆ωt)] e − γ rt + C3 e −2 γ r t , (5.58)
2∆ω − ( γ r /γ 0 )∆ω m
C1 = − ,
2∆ω − ∆ω m
∆ω(∆ω + ∆ω m ) + γ e2 − γ 02
C2 = ,
γ 0 (2∆ω − ∆ω m )
( γ /γ )∆ω m
C3 = e 0 .
2∆ω − ∆ω m 

In the above equation, the LS DC gain GLS is given by

2µ 2 γ 0 (2∆ω − ∆ω m )  ηω 0 
GLS = − Pin . (5.59)
(∆ω + γ r )[(∆ω + ∆ω m ) + γ r ]  nr 
2 2 2 2 

Note that since Δωm = −(ηω0/ng)V0, GLS depends on the applied volt-
age V0 as a result of the nonlinear transmission curve of the micror-
ing resonator. In the limit of a vanishingly small resonant frequency
shift (Δωm → 0), GLS reduces to the SS DC gain G in Equation 5.44.
The swing in the steady-state (or DC) transmitted power is given by
AC
∆Pout = Pout (t → ∞) = V0GLS. We also note that the LS response con-
tains a term which decays at a rate of 2γr. This fast-decaying term
comes from the second-order correction μ2|δam|2, which is neglected
in the SS analysis (see also footnote on page 261).
Figure 5.8 compares the transmitted power swings under the
LS condition and SS condition by plotting ΔPout/Pin versus the nor-
malized resonant frequency shift Δωm/γr for two sets of microring
parameters: the maximum gain case with γ = 1/2 and δ = 1/ 3 ,
and the critical coupling case with γ = 1 and δ = 1. As expected, the
SS transmission curves are linear for both cases but with differ-
ent gain slopes. The LS curves agree with the SS curves for small
frequency shifts (|Δωm/γr|< 0.1), but become highly nonlinear and
deviate from the SS curves at larger frequency shifts. We also
observe that the LS curves have asymmetric positive and negative
signal swings.
The plots in Figure 5.9 show the time-domain responses
AC
(Pout /Pin vs. γ r t) of an APMR modulator to step changes in the
resonant frequency of magnitudes Δωm/γr = ±0.1 and Δωm/γr = ±0.2.
The microring is assumed to be critically coupled (γ = 1) and the
­frequency detune δ is adjusted to show the cases of overdamped
(plot a), ­critically damped (plot b), and underdamped response
Active Photonic Applications of Microring Resonators 267

Steady-state transmitted power, ∆Pout/Pin


0.3
SS (δ = 0.58, γ = 0.5)
LS (δ = 0.58, γ = 0.5)
0.2
SS (δ = 1, γ = 1)
LS (δ = 1, γ = 1)
0.1

–0.1

–0.2

–0.3
–0.5 –0.4 –0.3 –0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5
Normalized frequency shift, ∆ωm/γr

Figure 5.8 Plot of the steady-state transmitted power swing, ΔPout/Pin,


versus the normalized resonant frequency shift, Δωm/γr, for two sets
of microring parameters: the maximum gain case with γ = 1/2 and
δ = 1/ 3 , and the critical coupling case with γ = 1, δ = 1. Solid lines are
LS solutions; dotted lines are SS solutions.

(plot c). Both LS responses (solid lines) and SS responses (dotted


lines) are shown for comparison, with the deviations between the
two solutions becoming more noticeable at larger step changes in
the resonant frequency. As expected, the time-domain responses
for the underdamped case are characterized by faster rise times
and larger overshoots than the overdamped case.

5.3.2 Large-signal response under sinusoidal modulation


When the microring resonant frequency is modulated by a large
sinusoidal RF signal, nonlinearity in the modulator response leads
to the generation of high-order harmonic frequencies which cause
distortion of the output modulated signal. We consider again
Equation 5.26 with an input monochromatic lightwave at f­ requency
ω and power Pin, sin = Pin e jωt . Suppose the resonant ­frequency of
the microring is modulated by a sinusoidal RF signal at frequency Ω,

δω m (t) = ∆ω m cos(Ωt) = ∆ω m (e jΩt + e − jΩt )/2, (5.60)

where Δωm is the amplitude of the resonant frequency shift.


Owing to the nonlinear transfer function of the modulator, new
harmonics are generated and mix with each other, so we expect
the wave ­amplitude in the microring to consist of an infinite series
of harmonics in general. However, for computational purposes,
268 Optical Microring Resonators
(a) (b)
0.15 0.15

Normalized power, Pout/Pin


Normalized power, Pout/Pin
0.1 ∆ω/γr = –0.2 0.1
0.05 ∆ω/γr = –0.1 0.05
0 0
–0.05 ∆ω/γr = 0.1 –0.05
–0.1 ∆ω/γr = 0.2 –0.1
–0.15 –0.15
0 2 4 6 8 10 0 2 4 6 8 10
Normalized time, γrt Normalized time, γrt
(c)
0.15
Normalized power, Pout/Pin

0.1
0.05
0
–0.05
–0.1
–0.15
0 2 4 6 8 10
Normalized time, γrt

Figure 5.9 Plots of the step response of an APMR modulator showing the
AC
normalized modulated power (Pout /Pin) versus the normalized time (γrt)
due to a step change in the resonant frequency: (a) overdamped response
with δ = 1/ 3 , γ = 1, (b) critically damped response with δ = 1; γ = 1, and
(c) underdamped response with δ = 1.2; γ = 1. The resonant frequency
shifts shown in each plot are Δωm/γr = ± 0.1 and ± 0.2. Solid lines are LS
solutions and dotted lines are SS solutions.

we will limit the series to the first N harmonics so that a(t) can be
expressed as
N

a(t) = ∑a e
n =− N
n
j( ω + nΩ )t
. (5.61)

Substituting the above expressions for δωm(t) and a(t) into Equation
5.26, with ω0 replaced by ω0 + δωm we get
N N
j∆ω m
∑ (Γ + jnΩ)a e
n =− N
n
jnΩt

2 ∑ a e
n =− N
n
j( n + 1)Ωt
+ e j( n −1)Ωt 
(5.62)
= (− jµ Pin ),
Active Photonic Applications of Microring Resonators 269

where Γ = jΔω + γr and δn0 is the Kronecker delta. By requiring that


the coefficients of terms of the same harmonic frequency satisfy the
equality in Equation 5.62, we obtain
j∆ω m
Γa0 − ( a1 + a−1 ) = − jµ Pin , (n = 0), (5.63)
2
j∆ω m
(Γ + jnΩ)an −
( an +1 + an −1 ) = 0, (n = ±1, ± 2, …± ( N − 1)),
2
(5.64)
j∆ω m
(Γ ± jNΩ)a∓ N − a( ± N −1) = 0, (n = ± N ). (5.65)
2
From the above expressions, we can solve for the amplitudes an to get

a0 = − jµ Pin F0 , (5.66)

j∆ω m
a± n = F± n a± ( n −1), (n = 1, 2,…N ),
2 (5.67)

where the terms F±n are computed backward starting from F±N,
1
F± N = , (5.68)
Γ ± jN Ω

1
F± n = , (n = 1, 2, …( N − 1)), (5.69)
Γ ± jnΩ + (∆ω m /2)2 F± ( n +1)

1
F0 = . (5.70)
Γ + (∆ω m /2)2 ( F1 + F−1 )

The transmitted power at the output of the modulator is


­determined from
N 2

∑a e
2 jnΩt
Pout (t) = sin − jµa = Pin − jµ n . (5.71)
n =− N

Expanding the above expression, we get


N N N

Pout (t) = Pin − jµ Pin ∑ (a e


n=− N
n
jnΩt ∗ − jnΩt
−a e
n )+ µ ∑ ∑ a a
2

n=− N n′=− N
∗ j( n− n′ )Ωt
n n′ e .

(5.72)
270 Optical Microring Resonators

If we write Pout(t) as a sum of harmonic frequencies,


N

Pout (t) = ∑Pe


n =− N
n
jnΩt
, (5.73)

then the power of the nth-order harmonic, Pn, can be obtained


by ­comparing Equations 5.72 and 5.73. Specifically, for the DC
­component we get
N

P0 = Pin − jµ Pin ( a0 − a0∗ ) + µ 2 ∑a


n =− N
n
2
, (5.74)

and for the nth-order harmonic,


N

Pn = − jµ Pin ( an − a ) + µ ∗
−n
2
∑a a
k =− N

k k −n . (5.75)

Note that the expression for the DC component can also be ­written
as
DC
P0 = Pout + µ2 ∑a
n≠0
n
2
, (5.76)

DC
where Pout is the transmitted DC power in the absence of a modulat-
ing signal (Equation 5.38), and the sum is over all n terms excluding
the DC term. Each of the terms |an|2 in the summation represents
the contribution to the DC power from the self-beating of each har-
monic, which is neglected in the SS analysis. The output AC power
of the APMR modulator is given by the first-order harmonic
N

P1 = − jµ Pin ( a1 − a ) + µ ∗
−1
2
∑a a
k =− N

k ( k −1) . (5.77)

Higher-order harmonic terms contribute to nonlinear effects. In


particular, the most dominant contribution is from the second-
order term
N

P2 = − jµ Pin ( a2 − a−∗2 ) + µ 2 ∑a a
k =− N

k ( k − 2) . (5.78)
Active Photonic Applications of Microring Resonators 271

However, in certain regimes of operation, the third-order harmonic


can also become nearly as large as the second-order harmonic, as
shown in the example below.
Figure 5.10a shows the normalized energies |an/A0|2 of the first
four harmonics (n = 0–3) in an APMR microring as a function of
the magnitude of the resonant frequency shift δm = Δωm/γr. The
­energies are normalized with respect to the energy in the microring
in the absence of a modulating signal, |A0|2 = |a0(0)|2. The micror-
ing has normalized frequency detune δ = 1 and critical ­coupling
(normalized coupling ratio γ = 1), and the normalized modulation
frequency is set at Ω = Ω/γ r = 0.1. A total of N = 10 ­harmonics are
used in the computation. The normalized transmitted powers at
the output of the modulator (Pn/Pin) of the first, second, and third

(a) (b)
0
Normalized power, Pn/Pin (dB)
Normalized energy, En/E0 (dB)

0
DC –10
–10 Ω
–20
–20 Ω
–30 –30 2Ω
–40 –40
2Ω
–50 –50
3Ω
–60 –60
–70 3Ω δ=1 –70 δ=1
–80 –80
0.01 0.1 1 0.01 0.1 1
Normalized frequency shift, ∆ωm/γr Normalized frequency shift, ∆ωm/γr

(c) (d)
0 0
Normalized power, Pn/Pin (dB)

Normalized power, Pn/Pin (dB)

–10 –10
Ω Ω
–20 –20
–30 –30
–40 –40 2Ω
2Ω
–50 –50
–60 –60 3Ω
3Ω
–70 δ = 0.5 –70 δ = 1.5
–80 –80
0.01 0.1 1 0.01 0.1 1
Normalized frequency shift, ∆ωm/γr Normalized frequency shift, ∆ωm/γr

Figure 5.10 (a) Plot of the normalized energies in the microring of the
first four harmonics as functions of the frequency shift. Plots of the output
modulator powers of the first three harmonics for (b) δ = 1, (c) δ = 0.5, and
(d) δ = 1.5. The APMR has critical coupling (γ = 1).
272 Optical Microring Resonators

log(Pn)
Ideal
ICP13 lines
CP ICP12
1dB

2Ω

3Ω
log(∆m)

Figure 5.11 Logarithm plot of output harmonic powers versus modula-


tion amplitude showing the 1 dB CP and intercept points ICP12, and ICP13.

harmonics are shown in Figure 5.10b. Figure 5.10c and d show the
output harmonics for frequency detune values of δ = 0.5 and 1.5.
We observe that the second harmonic dominates for all three cases
but the third harmonic can become as strong as the ­second har-
monic at smaller frequency detunes.
There are several parameters that can be used to quantify the
degree of nonlinearity of a modulator. One useful parameter is the
1 dB compression point (1 dB CP), which is defined as the mod-
ulation amplitude for which the output power of the first-order
­harmonic falls below the ideal straight line (or the SS response)
by 1 dB, as illustrated in Figure 5.11. Another useful parameter
is the i­ntercept point, which is used to quantify the significance
of the contributions of high-order harmonics to nonlinear distor-
tion. For example, to quantify the relative significance of the sec-
ond-order and third-order harmonics, we can denote ICP12 as the
intercept point of the ideal output powers (indicated by straight
dashed lines in Figure 5.11) of the first and second harmonics, and
ICP13 as the intercept point of the first and third harmonics. These
­parameters are also illustrated in Figure 5.11. In general, the smaller
the value of an intercept point, the more significant the contribu-
tion of the corresponding harmonic to nonlinear distortion.
Figure 5.12a and b plot the 1 dB CP and the intercept points as
functions of the frequency detune δ for an APMR modulator with
critical coupling (γ = 1). The normalized RF modulation frequency
is set at Ω = 0.1. We observe in Figure 5.12a that the 1 dB CP value
Active Photonic Applications of Microring Resonators 273
(a) (b)
1 dB compression point (δm) 2.5 16
14 ICP12
2.0

Intercept point (δm)


12
1.5 10
8
1 6
4
0.5
γ=1 2 ICP13 γ=1
0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Normalized frequency detune, δ Normalized frequency detune, δ

(c) (d)
1.2 5.5
1 dB compression point (δm)

5
Intercept point (δm)
ICP13
1.15
4.5
1.1 ICP12
4

1.05
3.5
δ=1 δ=1
1 3
0.1 1 10 0.1 1 10
Normalized coupling, γ Normalized coupling, γ

Figure 5.12 Plots of (a) the 1 dB CP and (b) intercept points versus the
normalized frequency detune δ for γ = 1 and Ω  = 0.1. Plots of (c) the 1 dB
CP and (d) intercept points versus the normalized coupling γ for δ = 1.

increases at larger frequency detunes, which implies that nonlin-


ear distortion is reduced. However, the relative dominance of the
second and third harmonics depends strongly on the frequency
detune. As Figure 5.12b shows, the ICP13 value is smaller than the
ICP12 value over the frequency detune range δ ~ 0.3–0.9 indicat-
ing that the third-harmonic component is the dominant source of
nonlinearity over this range of frequency detunes. Outside this
range, the second-harmonic component always dominates. Figure
5.12c and d show the 1 dB compression points and intercept points
as functions of the normalized coupling γ, for a fixed frequency
detune δ = 1 and normalized RF frequency Ω  = 0.1. We see that the
1 dB CP value is relatively independent of the normalized coupling,
and for all γ values, the second harmonic always dominates over the
third harmonic.
274 Optical Microring Resonators

5.3.3 Intermodulation products


The LS analysis in the previous section can also be extended to
the case where the microring is modulated by two RF sinusoi-
dal signals at two slightly different frequencies Ω1 and Ω2 (with
Ω1 − Ω2 ≪ Ω1, Ω2). Such an analysis is useful for evaluating the
second-order and third-order intermodulation products (IMP2
and IMP3), which are defined as the output powers at frequencies
Ω1 ± Ω2 and 2Ω1 − Ω2, respectively. The modulating signal is repre-
sented as a sum of two sinusoidal waves with frequencies Ω1 and Ω2
but the same amplitude Δωm:

δω m (t) = ∆ω m [cos(Ω1t) + cos(Ω 2t)] = ∆ω m (e jΩ1t + e − jΩ1t


+ e jΩ2t + e − jΩ2t )/2. (5.79)

The wave amplitude in the microring has a solution of the form


M N

a(t) = ∑ ∑a
m =− M n =− N
mn e j(ω + mΩ1 + nΩ2 )t , (5.80)

with a similar expression for the output power Pout(t). In Equation


5.80, M and N are the number of harmonics of Ω1 and Ω2 on each
side of the center frequency ω used in the analysis. By substitut-
ing Equations 5.79 and 5.80 into Equation 5.26 and following the
­procedure for the single-frequency analysis in the previous sec-
tion, we obtain the solutions for the wave amplitude amn and the
power Pmn associated with the harmonic frequency mΩ1 + nΩ2. For
a microring modulator, typical intermodulation products of interest
are the powers at the sum and difference frequencies Ω1 ± Ω2, and
the third-order intermodulation products 2Ω1 − Ω2 and 2Ω2 − Ω1.
Although the large number of h ­ armonic terms makes the computa-
tion somewhat more complex than the single-frequency analysis,
the intermodulation product analysis provides a more accurate way
to quantify nonlinear d ­ istortion in narrowband microring modula-
tors since the intermodulation ­frequencies are close to the input RF
frequencies Ω1 and Ω2. Finally, we note that the nonlinear distor-
tions of a microring modulator can be reduced by linearizing its
transfer function around the operating point. This approach has
been adopted to devise modulators with high linearity based on
Mach–Zehnder interferometers loaded with microring resonators
(Xie et al. 2003, Van et al. 2006).
Active Photonic Applications of Microring Resonators 275

5.4 Summary
Active photonic applications of microring resonators typically
involve electrically modifying the resonant frequencies of the
microrings. In this chapter, we reviewed the various physical mecha-
nisms that can be used to tune or modulate the resonant frequencies
of a microring, including the thermo-optic effect, the electro-optic
effect, and free carrier dispersion. Using the all-pass microring
modulator as a prototypical active microring device, we developed a
dynamic model based on energy coupling which allows us to study
the device behaviors in both the time and frequency domains and
under SS and LS conditions. These analyses also allow us to derive
expressions for important parameters characterizing the device per-
formance such as modulation efficiency, electrical bandwidth, har-
monic intercept points, and intermodulation products.

References
Agullo-Lopez, F., Cabrera, J. M., Agullo-Rueda, F. 1994. Electrooptics—
Phenomena, Materials and Applications. London: Academic Press.
Campbell, J. C., Blum, F. A., Shaw, D. W., Lawley, K. L. 1975. GaAs
electro-optic directional-coupler switch. Appl. Phys. Lett. 27(4):
202–205.
Cocorullo, G., Della Corte, F. G., Rendina, I. 1999. Temperature
dependence of the thermo-optic coefficient in crystalline
­silicon between room temperature and 550 K at the wavelength
of 1523 nm. Appl. Phys. Lett. 74(22): 3338–3340.
Dong, P., Liao, S., Feng, D., Liang, H., Zheng, D., Shafiiha, R., Kung,
C.-C. et al. 2009. Low Vpp, ultralow-energy, compact, high-speed
silicon electro-optic modulator. Opt. Express 17(25): 22484–22490.
Dong, P., Qian, W., Liang, H., Shafiiha, R., Feng, N., Feng, D., Zheng,
X., Krishnamoorthy, A. V., Asghari, M. 2010a. Low power and
compact reconfigurable multiplexing devices based on silicon
microring resonators. Opt. Express 18(10): 9852–9858.
Dong, P., Shafiiha, R., Liao, S., Liang, H., Feng, N., Feng, D., Li,
G., Zheng, X., Krishnamoorthy, A. V., Asghari, M. 2010b.
Wavelength-tunable silicon microring modulator. Opt. Express
18(11): 10941–10946.
276 Optical Microring Resonators

Leviton, D. B., Frey, B. J. 2006. Temperature-dependent absolute


refractive index measurements of synthetic fused silica. In SPIE
Astronomical Telescopes + Instrumentation, Orlando, FL, 62732K.
Li, C., Zhou, L., Poon, A. W. 2007. Silicon microring carrier-­injection-
based modulators/switches with tunable extinction ratios
and OR-logic switching by using waveguide cross-coupling.
Opt. Express 15(8): 5069–5076.
Liu, A., Jones, R., Liao, L., Samara-Rubio, D., Rubin, D., Cohen, O.,
Nicolaescu, R., Paniccia, M. 2004. A high-speed silicon optical
modulator based on a metal-oxide-semiconductor capacitor.
Nature 427(12): 615–618.
Milivojevic, B., Raabe, C., Shastri, A., Webster, M., Metz, P., Sunder,
S., Chattin, B., Wiese, S., Dama, B., Shastri, K. 2013. 112Gb/s
DP-QPSK transmission over 2427 km SSMF using small-
size silicon photonic IQ modulator and low-power CMOS
driver. In Optical Fiber Communication Conference, Anaheim, CA,
paper OTh1D.1.
Oh, M.-C., Zhang, H., Szep, A., Chuyanov, V., Steier, W. H., Zhang, C.,
Dalton, L. R., Erlig, H., Tsap, B., Fetterman, H. R. 2000. Electro-
optic polymer modulators for 1.55 μm wavelength using phen-
yltetrane bridged chromophore in polycarbonate. Appl. Phys.
Lett. 76(24): 3525–3527.
Pile, B., Taylor, G. 2014. Small-signal analysis of microring resonator
modulators. Opt. Express 22(12): 14913–14928.
Rabiei, P., Steier, W. H., Zhang, C., Dalton, L. R. 2002. Polymer
micro-ring filters and modulators. J. Lightwave Technol. 20(11):
1968–1975.
Raghunathan, V., Ye, W. N., Hu, J., Izuhara, T., Michel, J., Kimerling,
L. 2010. Athermal operation of silicon waveguides: spectral,
second order and footprint dependencies. Opt. Express 18(7):
17631–17639.
Sacher, W. D., Green, W. M. J., Assefa, S., Barwicz, T., Pan, H., Shank,
S. M., Vlasov, Y. A., Poon, J. S. K. 2013. Coupling modulation
of microrings at rates beyond the linewidth limit. Opt. Express
21(8): 9722–9733.
Sacher, W. D., Poon, J. K. S. 2008. Dynamics of microring resonator
modulators. Opt. Express 16(20): 15741–15753.
Active Photonic Applications of Microring Resonators 277

Sherwood-Droz, N., Wang, H., Chen, L., Lee, B. G., Biberman, A.,
Bergman, K., Lipson, M. 2008. Optical 4x4 hitless silicon router
for optical Networks-on-Chip (NoC). Opt. Express 16(20):
15915–15922.
Sze, S. M. 1981. Physics of Semiconductor Devices, 2nd Ed. New York:
John Wiley & Sons.
Van, V., Herman, W. N., Ho, P.-T. 2006. Linearized microring-loaded
Mach-Zehnder modulator with RF gain. J. Lightwave Technol.
24(4): 1850–1854.
Xie, X., Khurgin, J., Kang, J., Chow, F.-S. 2003. Linearized Mach–
Zehnder intensity modulator. IEEE Photonics Technol. Lett. 15(4):
531–533.
Xu, Q., Schmidt, B., Pradhan, S., Lipson, M. 2005. Micrometre-scale
silicon electro-optic modulator. Nature 435(19): 325–327.
Yariv, A. 1991. Optical Electronics, 4th Ed. Philadelphia: Saunders
College Publishing.

You might also like