optical-microring-resonators-theory-techniques-and-applications_compress
optical-microring-resonators-theory-techniques-and-applications_compress
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
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
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.
Preface ix
v
vi Optical Microring Resonators
Index 279
Preface
ix
x Optical Microring Resonators
V. Van
University of Alberta
September 2016
Chapter 1
Elements of an Optical
Microring Resonator
1
2 Optical Microring Resonators
(a) (b)
Bus waveguide Microring Bus waveguide Microdisk
n12 − n22 n1 − n2
∆n = ≈ . (1.1)
2n12 n1
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)
H( x , y , z) = H( x , y )e − jβz , (1.3)
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)
∇ × E = − jωµ 0 H, (1.6)
∇ × ∇ × E = n2 k 2 E , (1.8)
∇ 2E + n2 k 2E = ∇(∇ ⋅ E ). (1.9)
∂E
∇ 2t E t + (n2 k 2 − β 2 )E t = ∇ t (∇ t ⋅ E t ) + ∇ t z , (1.10)
∂z
∂Ez
∇ ⋅ ( n 2 E) = ∇ t ⋅ ( n 2 E t ) + n 2 = 0, (1.11)
∂z
∂Ez 1
= − 2 ∇ t ⋅ (n2E t ). (1.12)
∂z n
1
∇ 2t E t + (n2 k 2 − β 2 )E t = ∇ t (∇ t ⋅ E t ) − ∇ t 2 ∇ t ⋅ (n 2E t ) . (1.13)
n
Pxx Pxy Ex Ex
P = β2 , (1.14)
yx Pyy Ey Ey
∂ 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
y (µm)
y (µm)
y (µm)
y (µm)
0.2
y (µm)
0.2 0.2
0 0 0
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
1
∇ 2t Ht + (n2 k 2 − β 2 )Ht = − 2
(∇ t n2 ) × (∇ t × Ht ). (1.23)
n
Elements of an Optical Microring Resonator 9
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
* 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
y-slab I II I II II I II
neff(x) neff(x)
x-slab
neff,1 neff,1
neff,2 neff,2 n3 n3
x x
Ex ( x , y ) = X( x)Y( y ). (1.29)
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)
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)
d2 Hy
+ neff
2
( x)k 2 − β 2 H y ( x) = 0. (1.33)
dx 2
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.
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
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
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
dτ g (∆λ )2 d 2 τ g
τ g (λ ) = τ g (λ 0 ) + ∆λ + + (1.38)
dλ 2 dλ 2
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
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)
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)
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)
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.
∂2 F ∂2 F
2
+ 2 + n 2 ( x , y )k 2 F = 0, (1.45)
∂x ∂y
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
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
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
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
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
∫
Rrc
ψ(r ) dr
Reff = 0
. (1.57)
1 2
∫
Rrc
r ψ ( r ) dr
0
H z , quasi-TE mode,
F( r , θ ) =
Ez , quasi-TM mode.
24 Optical Microring Resonators
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
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
C1 J m (n1kr ), r ≤ R,
ψ(r ) = ( 2) (1.64)
C2 H m (n2 kr ), r > R,
J m (n1kR)
C2 = C1 . (1.65)
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)
* 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
2mω mn ′′
α mn = . (1.72)
ω ′mn Reff
′ t)e −2ω mn
U(t) = U 0 cos 2 (ω mn ′′ t
, (1.74)
1
U = U 0 e −2ω mn
′′ t
. (1.75)
2
dU
PL = − = 2ω ′′mn U , (1.76)
dt
Elements of an Optical Microring Resonator 27
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
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)
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
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)
104
TE
102
TM
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.
n1 n2
θ n2 n1 n2
0 0 x
n3 R1 R2 r n3
s
∇ 2E + n2 k 2E = ∇(∇ ⋅ E ), (1.82)
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
r → x + R, θ → s/R, z → y , (1.84)
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
1
∇ 2H + n2 k 2H = − 2
∇n2 × (∇ × H ), (1.91)
n
Elements of an Optical Microring Resonator 33
1 ∂ 2 Ht 1
∇ 2t Ht + 2 2
+ n2 k 2Ht = − 2 (∇ t n2 ) × (∇ t × Ht ). (1.92)
r ∂θ n
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
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
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
∇ × 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)
∇ ⋅ (E a × Hb − Eb × Ha ) = − jω(ε b − ε a )E a ⋅ Eb . (1.103)
∂
∂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
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
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
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:
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
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}.
da1 da2
C + = j(K12 + Cβ 2 )a1 ( z) + j(K 22 + β 2 )a2 ( z), (1.115)
dz dz
d a1 ( z) a1 ( z)
C = − j(K + BC) , (1.116)
dz a2 ( z) a2 ( z )
where
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
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)
k12
T12 = − j sin(Ωz)e − jγ 0 z , (1.129)
Ω
k 21
T21 = − j sin(Ωz)e − jγ 0 z . (1.130)
Ω
Uncoupled
Distributed coupler waveguides Lumped coupler
τ γ0
β1 γ0
–jκ –jκ
β2 γ0 γ0
τ
L L/2 L/2
(a) (b)
x x
R1 R2 R2
θ
s0 s0 s(θ)
dz
z z
0 0
κ= ∫
− π /2
kc (s)R2 cos θdθ. (1.136)
(a) (b)
0.4 0.4
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.
Waveguide core
(c) (d) Over-cladding layer
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
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
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
Analytical Models of a
Microring Resonator
53
54 Optical Microring Resonators
2πnr R
λm = . (2.2)
m
dφrt dnr nr 2π
= − 2πR. (2.3)
dλ dλ λ λ
dφrt 2π
= − ng 2 2πR. (2.4)
dλ λ
2π λ2
∆λ FSR = =− . (2.5)
dφrt /dλ 2πng R
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)
ω m α/2 cα
γm = = . (2.8)
m/Reff 2nr
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 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
2πnr
Q0 = . (2.12)
αλ m
(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
Ei (r , t) = E0 e( x , y )e j(ωt −β b z ) , (2.13)
st = − jκ 1D + τ1si , (2.14)
* 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
sd = − jκ 2 B, (2.16)
C = τ 2 B. (2.17)
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
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
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
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
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
0
∆φBW Insertion
Transmission (dB)
–5 loss
–10 Td
–15 Tt
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
Td ,max
Td (φrt ) = , (2.31)
1 + F sin 2 (φrt /2)
4τ1τ 2 art
F= . (2.33)
(1 − τ1τ 2 art )2
Td ,max T
Td (∆φ BW /2) = = d ,max , (2.34)
1 + F sin 2 (∆φ BW /4) 2
which gives
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
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
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,
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
4τart
F= , (2.46)
(1 − τart )2
2
τ − art (2.47)
Tap,min = .
1 − τart
(a) (b)
τ τ=1
R
si Perfectly
B A reflecting
st mirror
κ, τ
Input port Output port
si st
(a) (b)
1 12
Critically coupled
Power enhancement (|FE|2)
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
λ 2m 1 − τart
∆λ BW = . (2.48)
2π 2 ng R τart
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
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 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
( p02 − 1)Trt /2
τ g ,d (φrt ) = . (2.55)
1 + p02 − 2 p0 cos φrt
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
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,
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)
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
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
dψ
GD × BW =
∫
FSR
τ g,d dω = −
∫
FSR
dω
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)
–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.
dτ g d2 ψ
Dg = ≈ −Trt2 2 , (2.66)
dω dφrt
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)
* 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
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).
da
= ( jω 0 − γ 0 − γ 1 − γ 2 )a. (2.69)
dt
da
= ( jω 0 − γ 0 − γ 1 − γ 2 )a − jµ1si . (2.70)
dt
µ12 = 2γ 1 . (2.73)
2 2 2 2πR
a(t) = A(t) Trt = A(t) , (2.74)
vg
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
− jµ1si
a= , (2.78)
j(ω − ω 0 ) + γ
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
− 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 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
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
1 1 1 1
= + + . (2.88)
Q Q0 Q1 Q2
Analytical Models of a Microring Resonator 81
ω0 2π 2 ng R
Q= = , (2.89)
2µ 2 λ 0κ 2
s + γ − µ 2 s + γ 0 − µ 2 /2
H ap (s) = = . (2.90)
s+ γ s + γ 0 + µ 2 /2
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
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
− jκ 1si
A= . (2.95)
1 − τ1τ 2 art e − jφrt
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
−κ 2 µ2
ln(τ1 ) =
1
2
( 2
)
ln 1 − κ 12 ≈ 1 = − 1 Trt ,
2
(2.99)
µ2 µ2
1 − τ1τ 2 art e − jφrt ≈ 1 + 2 + ( γ 0 + j∆ω) Trt , (2.100)
2 2
− jµ1si
a= . (2.102)
( j∆ω + γ 0 ) + µ12 /2 + µ 22 /2
–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
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
87
88 Optical Microring Resonators
T T T
bk bk+1
Λ
ak +1 ak − jβΛ
b = b e , (3.2)
k +1 k
1
cos(βΛ) = (T11 + T22 ). (3.5)
2
k−1 k k+1
κ κ κ
Λ Λ
ak ck ak+1 ck+1
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
jφ
H ap = = − e , (3.9)
1 − τe − jφ 1 − τe − jφ
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
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
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
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
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
κ κ
ck–1 ak dk bk+1
dk–1 bk ck ak+1
k−1 k k+1
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
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
∏ (s − z ) ,
M
P(s) K 0 k =1
k
H(s) = = (3.24)
∏ (s − p )
N
Q(s)
k
k =1
* 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
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
R1 R2 R3 RN
κ1 κ2 κ3 κN
si st
∏
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)
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
∏ (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
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.
(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
µ0 µ1 µi µN–1 µN
si
ai
1 2 i i+1 N–1 N
st sd
Through Drop
vg κi
µi = κ i = , i = 0, N , (3.39)
2πRi Trt , i
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
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
ω − ω i = (ω − ω 0 ) − (ω i − ω 0 ) = ∆ω − δω i , (3.43)
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
(− j)N +1 (µ 0 µ1µ 2 µ N )
Hd (s) = , (3.49)
CN (s)
* 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
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.
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)
1 1
a = sinh sinh −1 ,
N ε
(3.60)
1 1
b = cosh sinh −1 .
N ε
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 ]
kπ 1 1 + ε 2 + 1
ck2 = sin 2 + sinh 2 ln , 0 ≤ k ≤ N (3.62)
N 2N 1 + ε 2 − 1
0
Transmission (dB)
–10
–20
Drop
–30
–40 Through
–50
–100 –50 0 50 100
Frequency detune, ∆f (GHz)
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
that the filter passband becomes flatter and the roll-off becomes
steeper as the filter order is increased.
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
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
ai +1 1/2 ai
b = z K iP b . (3.65)
i +1 i
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
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
z − N /2
Hd ( z) = . (3.69)
T12
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 )
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
( 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.
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
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)
∆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)
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
–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
ck ak
d = Mk b . (3.85)
k k
(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
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
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
Pk = −κ 2k z −2 Rk −1 + (1 − τ 2k z −1 )Pk −1 , (3.96)
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.
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
2 2k 2N
κ1, 1 L κ1, k L κ1, N
Drop
bk dk bk + 1 sd
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.
1 Fk ( z ) jK k z −1
−1
Mk = , (3.101)
Gk ( z −1 ) jK k z −1 Fk ( z −1 )
K k = κ 12, k κ 2 , k , (3.102)
γ k 0
Pk = e − jβ b L , (3.105)
0 1
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
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
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
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
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)
si µi µ12 µ23
1 2 3 m
st
µ1, 2m
N N–1 N–2
sd
µo
(a) (b)
1 2
3
4 3 1 2
ω − ω i = (ω − ω 0 ) − (ω i − ω 0 ) = ∆ω − δω i , (3.117)
L = diag[γ i , 0, … 0, γ o ], (3.119)
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
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,
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
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
Λ )−1 W T u.
a = W(sI + jΛ (3.131)
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
N
a WN , kW1, k
AN 1 = N
u1 µo = 0
= ∑
k =1
s + jλ k
. (3.135)
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
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
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
∑W
k =1
2
1, k = 1. (3.151)
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
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
(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)
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 0.8209 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)
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]
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
K12
K 34
M1 = , (3.163)
Ki, j
τ i , j − jκ i , j
Ki, j = . (3.164)
− jκ i , j τ i , j
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 = La′ + s, (3.166)
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 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)
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)
κ iκ o N
−
τo
∑ QN,kQk−,11 = 0, (3.176)
k =1
z −1P( z −1 )
Hd ( z −1 ) = , (3.177)
Q( z −1 )
* 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)
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
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
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
163
164 Optical Microring Resonators
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
(4.3)
XPM
PNL (ω1 ) = 6ε 0 χ( 3 ) |E(ω 2 )|2 E(ω1 ), (4.5)
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)
n = n0 + n2 I , (4.8)
166 Optical Microring Resonators
Dχ( 3 )
n2 = (4.12)
4n02 ε 0 c
n = n0 + n2 〈E 2 (t)〉. (4.13)
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.
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
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
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):
∂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
1 ∂2E ∂ 2 PL ∂ 2 PNL
∇2E − = µ 0 + µ 0 , (4.25)
c 2 ∂t 2 ∂t 2 ∂t 2
* 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
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
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
∇ 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
∆β NL =
k2 ∫εC
r,NL |φ( x , y )|2 dx dy
. (4.39)
2β
∫|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
∂A
+ j(β − β0 )A = − j∆β NL A. (4.41)
∂z
∂A α
+ j[β(ω) − β0 ]A + 0 A = − j∆β NL A. (4.42)
∂z 2
∂A α
+ j(∆ωβ1 + 12 ∆ω 2β 2 + )A + 0 A = − j∆β NL A, (4.43)
∂z 2
∂A jβ 2 ∂ 2 A
∂A α0
+ β1 − + = − j∆β NL A
A , (4.44)
∂z ∂t 2 ∂t 2 2
∆β NL =
3 χ( 3 ) k 2 ∫ |φ(x, y)| dx dy |A(z, ω − ω )| .
C
4
0
2
(4.45)
2β
∫|φ(x, y)| dx dy
2
∆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
∫ ∫
( 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
∂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
3 χ( 3 ) ω
γ= 2
, (4.49)
4neff ε 0 c 2 Aeff
(∫|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
α0
∂A
+ = − j γ |A
A |2 A
. (4.52)
∂z 2
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,
Substituting the above result into Equation 4.54 and solving for the
nonlinear phase φ, we get
1 − e −α0 z
zeff = . (4.57)
α0
( z , t) = A
A (0, t)e − α0 z/2 e jϕ ( z ,t ) , (4.58)
A( z) = A0 e − α0 z/2 e jϕ ( z ) , (4.60)
1 − e −2α0 πR a2 − 1
Leff = = 2πR rt . (4.61)
α0 2 ln art
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.
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)
∆β fc =
k2 ∫εC
r,fc |φ( x , y )|2 dx dy
=
nc k
(σ r − jσ a /2k )N fc , (4.66)
2β
∫ neff
2
|φ( x , y )| dx dy
N ( z , t) =
fc
∫ N (r, t)|φ(x, y)| dx dy .
C
fc
2
(4.67)
∫|φ( x , y )|2 dx dy
nc2 k jα 2
γK = 2 n2 − 2k , (4.69)
neff Aeff
nc k jσ a
γ fc = σr − . (4.70)
neff 2k
∂N fc 1 1
+ + N fc = G, (4.71)
∂t τ rec τ diff
∫ D∇ N
C
2
fc |φ( x , y )|2 dx dy
=−
N fc
. (4.72)
∫
|φ( x , y )| dx dy τ diff
2
G=−
1 ∂I
=−
1 ∫ (∂I/∂z)|φ(x, y)| dx dy .
C
2
(4.73)
2ω ∂z 2ω
∫
|φ( x , y )|2 dx dy
we have
∂I ∂U
= 2neff ε 0 c|φ( x , y )|2 2U , (4.75)
∂z ∂z
∂U
= Im{∆β K }U , (4.76)
∂z
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
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
∂N fc N fc ( z , t)|4 ,
+ = ζ TPA |A (4.80)
∂t τ fc
1 1 1
= + . (4.81)
τ fc τ rec τ diff
180 Optical Microring Resonators
α FCA
max n σ a Ep
ra = TPA
= c , (4.82)
α max neff 2 2ωAeff
∂(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
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)
dN fc N fc
+ = ζ TPA P 2 ( z , t). (4.88)
dt τ fc
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)
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)
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
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
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),
dAi − jβi z
− j 2βi φi ( x , y ) e = −µ 0 ω i2 PNL (r , ω i ), (4.98)
dz
∗
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 , 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)
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)
Ai
Ai → ,
( )
1/2
2ni ε 0 c |φi |2
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
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
dA1
dz
(
= − jγ |A1|2 + 2|A2|2 A1 , ) (4.115)
dA2
dz
(
= − jγ |A2|2 + 2|A1|2 A2 , ) (4.116)
A1 ( z) = P1 e − jγ ( P1 + 2 P2 )z , (4.117)
A2 ( z) = P2 e − jγ ( P2 + 2 P1 )z , (4.118)
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
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
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 Pp = |Ap(0)|2 is the input pump power. The solutions for the
signal and idler waves can be expressed as
where
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
γPp
Ai ( z) = j Ps sinh( gz)e − jΩz . (4.139)
g
A(z)
z=0
0 –φNL
dA α 0
+ A = − jγ |A|2 A, (4.141)
dz 2
− jκ Pin
A(0) = . (4.144)
1 − τart e j( φL + φNL )
κ 2 Pin
Pr,max = . (4.146)
(1 − τart )2
{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
∆ω NL n φ n γP
= − r NL ≈ − r 0 , (4.149)
ω n
g Lφ ng βr
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)
− jκ Pin
A0 = 2 , (4.151)
1 − τart e jφL e − jξ|A0|
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)] ,
{ } 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
The above equation describes the discrete time evolution of the per-
turbation ε(t). Writing the solution as
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)
ωm ωm+1
Low
power
ω
High
FSR/2
power
ωm ω0 ωm+1 ω
(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
∆nNL
∆ω NL = −ω 0 , (4.162)
ng
Nonlinear Optics Applications of Microring Resonators 203
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
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
2δ + δ 2 − 3 2δ − δ 2 − 3 (4.175)
− < E0 < − .
3h 3h
−Γ − jηK′ a02
M= ∗ 2 , (4.176)
jηK′ ( a0 ) −Γ ∗
−Γ 0 − jηfc′ a0
M= 0 −Γ ∗ jηfc′ a0∗ , (4.180)
2 ∗ 2
2ς a0 a0 2ς a0 a0 −1/τ fc
λ 3n + a2 λ n2 + a1λ n + a0 = 0, (4.181)
Nonlinear Optics Applications of Microring Resonators 207
a2 = 2 + 1/τ n ,
2
a1 = 1 + 2/τ n + δ NL ,
2
a0 = (1 + δ NL + 4 hE02δ NL )/τ n .
∆ = (δ 2 + 3) − (τ 2n − 2/τ n ). (4.183)
* 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
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
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.
(b) (c)
0.25 0.12
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 .
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 4.9b 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
dT
= γks L, (4.188)
dPp max
(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
dTap 2τ 2κ 4 sin φ
= . (4.191)
dφ (1 + τ 4 − 2τ 2 cos φ)2
dTap 3 3
≈ . (4.192)
dφ φmax
8κ 2
κ 2 Pin,p
Pr,p = , (4.194)
1 + τ 4 − 2τ 2 cos φ
dTap 9 3 γks L 4
≈ ∝ FE , (4.195)
dPin,p max
8 κ4
Nonlinear Optics Applications of Microring Resonators 217
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
∂A α 0
+ A = − jγ |A|2 A, (4.197)
∂z 2
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 )
− 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)
(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
(c) (d)
1
0.6
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
(a) 1 (b) 30
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
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
− 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
Normalized power
0.3 Pump 0.6
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
∂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 0
–3 –2 –1 0 1 2 3 –3 –2 –1 0 1 2 3
Time t/T0 Time t/T0
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
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)
j 2 γPp zeff
Ai∗ ( z) = Bi∗ ( z)e − α0 z/2 e , (4.218)
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
Ai∗ ( z) = [ai∗ cosh( gz) + bi∗ sinh( gz)]e − α0 z/2 e jΩ( z ) , (4.222)
where
K = ∆β + 2γPp . (4.225)
At the coupling junction, the signal and idler waves satisfy 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
jγPp
ai∗ = as = f 2 as , (4.231)
gf1 − jK/2
− 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 ) ,
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 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
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:
(a) (b)
0.32 101
Pin,p = 250 mW
Conversion efficiency, ηr
Conversion efficiency, ηr
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
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
mi c
δω i = ω i − Ωi = (2ω p − ω s ) − , (4.240)
nr (Ωi ) R
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 )
2D1∆ω 2
δω i ≈ . (4.246)
ng − D1∆ω
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
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
241
242 Optical Microring Resonators
dneff ∂n ∂n
(2πR − L) ∆λ + L eff ∆λ + eff V = m∆λ , (5.2)
dλ ∂λ ∂V
dn n ∂n
2πR eff − eff ∆λ + L eff V = 0, (5.3)
dλ λm ∂V
λm L
∆λ = ∆neff (V ), (5.4)
ng 2πR
* 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
∂T
ρC = ∇ ⋅ ( K ∇T ) + q , (5.8)
∂t
Micro-heater y
SiO2 cladding
+V –V Si core
x
SiO2 buffer
Si substrate
ηP λ m
∆λ = . (5.9)
2πR ng
* 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
(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)
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
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 2 1 2 1 2
x + 2 y + 2 z + 2r41Ex yz + 2r41Ey zx + 2r41Ez xy = 1, (5.15)
n02 n0 n0
1 2 1 2 1 2
n2 x + n2 y + n2 z + 2r41Ex yz = 1. (5.16)
0 0 0
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
1 3
nY ≈ n0 + n0 r41Ex , (5.18)
2
1
nZ ≈ n0 − n03 r41Ex . (5.19)
2
250 Optical Microring Resonators
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
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
∗
Si substrate Si substrate
(c)
Contact Contact
SiO2 p-Si
SiO2
Si substrate
Pout(t)
vm(t)
Pin Pout(t)
ω0 ω
δωm(t)
dδω m δω ηω 0
= − m − vm (t). (5.28)
dt τ m ng τ m
− jµ Pin − jµ Pin
a0 = = , (5.29)
j∆ω + γ r Γ
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
ηω 0 Vm ( jΩ)
δω m ( jΩ) = − , (5.32)
ng τ m jΩ + 1/τ m
ηω 0 ja0Vm ( jΩ)
δam ( jΩ) = − . (5.33)
ng τ m ( jΩ + Γ )( jΩ + 1/τ m )
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)
AC
µ2 ∗ µ2
Pout (t) = jµ Pin 1 − δam (t) − 1 − ∗ δam (t) . (5.40)
Γ Γ
AC
µ2 ∗ µ2
Pout ( jΩ) = jµ Pin 1 − δam (− jΩ) − 1 − ∗ δam ( jΩ) . (5.41)
Γ Γ
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
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
ηP δ γ
G = −16Q0 in 2 2 3
, (5.45)
ng (1 + δ ) (1 + γ )
Active Photonic Applications of Microring Resonators 259
4Q0 ηPin
Gmax = − . (5.46)
3 3 ng
G 12 3δ γ
= , (5.47)
Gmax (1 + δ 2 )2 (1 + γ )3
–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
2
GΓ ( jΩ + 2γ 0 )
H m ( jΩ) = . (5.49)
2γ 0 ( jΩ + Γ )( jΩ + Γ ∗ )
Active Photonic Applications of Microring Resonators 261
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
–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
~
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
–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.
ja0 ∆ω m
δam (t) = (1 − e − Γt ). (5.52)
Γ
* 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.
a new steady state due to the resonant frequency shift Δωm. Setting
daf/dt = 0 in Equation 5.27, we get
− jµ Pin
af = . (5.54)
Γ − j∆ω m
dan
= −Γan ,
dt
which yields
an (t) = A0 e − Γt . (5.55)
jµ Pin j∆ω m − Γt
a(t) = e − 1 . (5.56)
Γ − j∆ω m Γ
∆ω m µ Pin
∆am = .
Γ(Γ − j∆ω m )
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
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
–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
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
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 )
∑a e
2 jnΩt
Pout (t) = sin − jµa = Pin − jµ n . (5.71)
n =− N
n=− N n′=− N
∗ j( n− n′ )Ωt
n n′ e .
(5.72)
270 Optical Microring Resonators
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)
P2 = − jµ Pin ( a2 − a−∗2 ) + µ 2 ∑a a
k =− N
∗
k ( k − 2) . (5.78)
Active Photonic Applications of Microring Resonators 271
(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)
–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
Ω
2Ω
3Ω
log(∆m)
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
(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.
a(t) = ∑ ∑a
m =− M n =− N
mn e j(ω + mΩ1 + nΩ2 )t , (5.80)
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
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.