Você está na página 1de 7

2206

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 25, NO. 8, AUGUST 2007

Multiwavelength Transmission Microcavity in SOI


Planar Ridge Waveguides
Xiaohua Shi, Wei Ding, and Duncan W. E. Allsopp, Member, IEEE

AbstractA novel waveguide-based 1-D photonic crystal (PhC)


microcavity that simultaneously transmits light within the 1310and 1550-nm wavelength telecommunications bands has been
realized. A novel semianalytical model based on wave-vector diagrams was developed to enable rapid optimization of the 1-D
PhC with complex periodicity to be etched into an SOI ridge
waveguide. Measurements of the multiwavelength transmission
characteristics of the resulting complex PhC microcavity revealed
transmission peaks at 1321 and 1562 nm with nearly equal transmittivity that is close to the design specification.
Index TermsGuided waves, microcavity, multiwavelength
filter, photonic crystal (PhC) slabs.

I. I NTRODUCTION

URING the last two decades, photonic crystal (PhC)


structures have received considerable attention in the optics community [1][3]. The periodic variations in the dielectric
constant in a PhC give rise to novel optical properties that
depend on the period, the dielectric contrast, and the geometry
of each period [4]. In particular, the application of PhCs to the
formation of resonant microcavities has been widely studied
with reports of both 1-D [5][9] and 2-D structures with ultrasmall volume and high quality factor (Q) [10], [11].
The resonant wavelengths of a microcavity are those obeying
the standing wave condition, which are set by the width of the
central layer sandwiched between the PhC reflectors, leaving
little or no scope for engineering the spacing of the transmission
peaks. Recently, Lee et al. demonstrated a novel multiresonance
microcavity [12]. The structure comprised of multiple layers of
Si-SiO2 with the homogeneous in-cavity material replaced by a
Si-SiO2 1-D PhC of different period to that of the reflectors.
The second PhC introduces a degree of flexibility into the
design of the microcavity response, opening the prospect of
creating a small volume filter with narrow passbands: one for
transmitting system control information and the other for client
data transmitted in distinct waveguide bands.
Multiwavelength filters are desirable in a number of application areas, notably optical fiber communications networks
in which system control information and data are transmitted
simultaneously over the same fiber, but encoded on carriers of
Manuscript received November 16, 2006; revised March 13, 2007. This work
was supported in part by the U.K. universities Overseas Research Student
(ORS) scheme and in part by Photeon GmbH of Austria.
X. Shi and D. W. E. Allsopp are with the Department of Electronic and
Electrical Engineering, University of Bath, BA2 7AY Bath, U.K.
W. Ding is with the Department of Physics, University of Bath, BA2 7AY
Bath, U.K.
Digital Object Identifier 10.1109/JLT.2007.901838

different wavelength. The architecture of such systems, which


is allied with the drive toward photonic component integration,
means that guided wave device geometry is more convenient
planar PhC filter structures. However, realizing the multiwavelength filters based on the dual PhC concept in a waveguide
configuration presents a number of challenges. The filter passbands should be narrow and ideally with equal transmittivity at
strategic wavelengths, for example, in the 1310- and 1550-nm
communications windows for applications in optical networks.
Replacing a solid cavity layer with a PhC will inevitably
increase scattering losses, which ideally should be the same for
all transmitted bands.
This paper describes an investigation of multiwavelength
transmission in 1-D PhC microcavities formed in ridge
waveguides, with the aim of realizing compact photonic
integrated circuit filters capable of transmitting light simultaneously in the telecommunications bands centered on 1310and 1550-nm wavelengths. The use of a second PhC in the
microcavity introduces complexity into the design process, for
which a simple model is desirable for rapid exploration of the
widened parameter space. The transfer matrix model, which is
applicable to complex planar dielectric filters and enables rapid
optimization, is not strictly valid for PhC structures etched into
waveguides, and an alternative approached is required.
First, an intuitive semianalytical model of multiwavelength
transmission in a 1-D PhC is developed. This is based on a
wave-vector diagram [13], and it provides a simple method
for rapid design optimization. The model can be applied to
both multilayer stack structures as well as 1-D PhCs etched or
imprinted into slab waveguides. The results of this new model
are compared with those of the transfer matrix method (TMM)
approach and of a more rigorous method that uses the guided
modes of a slab to describe a Bloch mode [14].
The new model is then applied to the realization of a
multiwavelength transmission microcavity formed in a siliconon-insulator (SOI) ridge waveguide. Next, the fabrication technology is described, and the measurements of the spectral
response of the complex microcavity filter are presented and
discussed in the context of the predictions of the novel semianalytical model.
II. T HEORY
Fig. 1 shows the different types of 1-D PhC microcavity
considered in the literature (see, for example, [12] and [14]).
Fig. 1(a) shows schematically a multiple layer 1-D PhC in
which the periodicity varies. The two Bragg reflectors at each

0733-8724/$25.00 2007 IEEE

SHI et al.: MULTIWAVELENGTH TRANSMISSION MICROCAVITY IN SOI PLANAR RIDGE WAVEGUIDES

Fig. 1.

2207

Schematic 1-D PhC microcavity structures. (a) Multilayer stacks. (b) PhC slabs. (c) Planar ridge waveguide on SOI wafers.

end of the microcavity consist of pairs of layers of high refractive index contrast, with period A , while the cavity layer itself
comprises a second PhC of period B , where A = B . In the
work described here, the layer pairs were taken to be silicon
(dielectric constant Si = 12) and air (air = 1) [15] for later
comparison with SiAir equivalent PhC structures etched into
SOI ridge waveguides.
Fig. 1(b) and (c) shows such equivalent structures formed in
a planar slab waveguide and a ridge waveguide. Since a ridge
waveguide forms a convenient means of coupling light in and
out of the microcavity and a basis for further device integration, the structure shown in Fig. 1(c) is of particular interest.
However, the field confinement in the x-direction associated
with the ridge waveguide width introduces a complication to
the analysis. While the field confinement in the x-direction is
likely to affect the coupling of waveguide modes with the Bloch
modes of the PhC device, within the latter, it has an insignificant effect compared with the confinement in the out-of-plane
y-direction associated with the slab thickness. Therefore, in the
theory that follows, the PhC structure is treated like that in
Fig. 1(b), in which the etched features are assumed to extend
laterally over the full range |x| .
Another difference between the structures shown in Fig. 1(b)
and (c) is that the air gaps in the latter do not extend through
the entire thickness of the Si layer. Incomplete etching of the Si
layer helps in controlling the radiation loss that occurs for larger
air fractions in air/Si PhCs [8]. On the other hand, the remaining
Si at the bottom of the air gaps should be thin enough to ensure
little effect on the design of a practical multiwavelength cavity.
The resonant wavelengths are the key factor in designing a
multiwavelength cavity with a PhC central layer of the type
shown in Fig. 1. For all three structures, the resonant condition
is given by
2s = 2(L + Lp )ks +
= 4(L + Lp )ns /s + ,

s = 1, 2, 3 . . . .

(1)

In (1), ks , s , and ns are, respectively, the propagation constant, the free space wavelength, and the effective modal index
of the sth resonant mode, L is the effective cavity length, and
Lp is the penetration depth of the resonant mode into the two
reflectors. With a PhC replacing the usual homogeneous region
in the central section of the cavity, a resonant mode comprises

Fig. 2. Wave-vector diagrams of 1-D PhC stacks with periods of A


(dot curves) and B (solid curves). B = 2A , Si = 12, and air = 1, with
equal optical length condition (nh lh = nl ll ) satisfied. The Bloch wave vectors
(kb4 , kb5 , and kb6 ) from (2) and the transmission frequencies (f4 , f5 , and f6 )
from TMM calculations are both indicated. The cavity length is twice B .

two counter-propagating Bloch waves. One originates from


reflections at the interfaces as light passes from high index
Si layers to low index air layers [leftward propagating Bloch
modes in Fig. 1(a)]. The other originates from reflections at the
interfaces as light passes from low index air layers to high index
Si layers [rightward propagating Bloch modes in Fig. 1(a)]. The
factor added to the right-hand side of (1) accounts for the
phase difference between the two types of reflection.
The use of Bloch wave-vector diagrams reveals the advantage of using a PhC in the central section of the cavity [5].
Fig. 2 shows the Bloch wave-vector diagrams for the two
1-D periodic stacks A (dotted lines) and B (solid lines), and
B = 2A . Both sets of bilayers satisfy the equal optical length
condition, nh lh = nl ll , where Nh,l and lh,l are, respectively,
the refractive index and thickness of the high (low) index layers.
Fig. 2 also shows by vertical dashed lines the Bloch wave
vectors (denoted as kB4 , kB5 , and kB6 ) of the fourth-, fifth-,
and sixth-order resonant modes calculated from (1), neglecting
the penetration depth (i.e., with Lp = 0).
In the case of a 1-D PhC microcavity of the type shown in
Fig. 1(a), the transmission frequencies can be calculated using
TMM. The corresponding values are also shown as horizontal
dashed lines in Fig. 2 and are labeled as f4 , f5 , and f6 . If the
two methods used in calculating the resonant condition were

2208

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 25, NO. 8, AUGUST 2007

Fig. 3. Schematic illustration of a plane wave and its component wave vectors
propagating via total internal reflection through a slab waveguide of permittivity
av , which is equivalent to the 1-D PhC.

entirely equivalent, the vertical and horizontal dashed lines


would intersect on the appropriate branches of the wave-vector
diagrams. In this event, a frequency mismatch of less than 3% is
obtained, which can be ascribed to the neglect of the penetration
depth in (1). This demonstrates that (1) and the TMM can be
used interchangeably in calculating the resonant wavelengths
of a complex multiple-layer 1-D PhC microcavity of the type
shown in Fig. 1(a).
While the TMM additionally provides and rigorously calculates the band gaps of any arbitrary multilayer 1-D PhC
microcavity, the wave-vector diagram approach enables rapid
optimization of the PhC central section of multiwavelength
transmission microcavities of the type shown in Fig. 1. The
wave-vector diagrams of the stack B can be tuned by varying
the air fraction and the refractive index contrast of the periodic
structure until its passband lies within the first band gap of the
1-D PhCs of period A used to form the reflectors. Since the
transmission wavelengths are adjustable with several degrees
of freedom, (1) provides a straightforward tool in designing a
multiwavelength transmission filter. Another advantage of the
wave-vector diagram model is its scope for application to more
complicated 1-D PhC slab waveguide structures.
So far, the discussion has focused on planar structures of the
type shown in Fig. 1(a). In contrast with the 1-D multilayer
stacks, the confined nature of the guided waves must be taken
into account when the 1-D PhCs are etched into the slab
waveguide structures of the type shown in Fig. 1(b). Fig. 3
illustrates the simple wave-vector model for a propagating
mode in a slab waveguide of dielectric constant equal to the
spatial average av over the 1-D PhC. The wave-vector diagram
gives the following dispersion relationship for such a mode:


c h2 (kz ) + kz2
c2 h2 (kz )
ck
=

+ f02 (kz ).
f (kz ) =

2 av
2 av
(2)2 av
(2)
In (2), c is the speed of light in free space, k is the wave
vector of the corresponding plane wave (see Fig. 3), kz is the
in-plane wave-vector amplitude (propagation constant), f (kz )
is the frequency, and h(kz ) is the vertical component of the
wave vector of a mode propagating inside the PhC with an inplane wave-vector amplitude (propagation constant) kz . The
term f0 (kz ) is the frequency of the in-plane cavity mode.
Equation (2) also describes the textbook dispersion relationship for a homogeneous slab waveguide of dielectric constant
equal to av .
It is proposed here that (2) can be adapted in a phenomenological way to account for the vertical confinement of the

Fig. 4. Calculated wave-vector curves of PhC silicon slabs using the approximate method of (2) (solid lines) and the rigorous method in [8] (dashed lines).
Only TE modes are calculated. The overlap layers are air. The substrate layers in
(a) and (b) are air and silica, respectively. The thickness of the slab is 0.4 times
the period; the air fraction is 0.3, with Si = 12, air = 1, and silica = 2.1.

propagating modes in the slab waveguide and for the action of


the 1-D PhC lattice in forming the Bloch waves in the in-plane
direction. If the vertical dielectric contrast is large, the outof-plane wave-vector components h(kz ) of the Bloch modes
inside the 1-D PhC will resemble the vertical field distribution
of the guided modes inside the equivalent slab waveguide of
dielectric constant equal to av [14]. This way, the relationship
between h(kz ) and kz can be well approximated by solving the
corresponding slab waveguide. On the other hand, the in-plane
components of the wave vectors of the guided waves kz interact
with the periodic lattice and give rise to band splitting when
kz approaches the edge of the Brillouin zone according to the
BlochFloquet theorem. This band splitting is expressed in the
frequency f0 (kz ), and the effect of the periodic lattice on kz
is calculated by using the plane-wave expansion method in an
infinite 1-D periodic multilayer.
The rationalization for (2) is that it accounts for the dominant
physics, albeit in a phenomenological way. The h(kz ) term
provides an information about the blue shift in the dispersion
relationship arising from the strong field confinement to the
equivalent slab waveguide [14]. The kz2 term contains the information about the large band splitting arising if there is a strong
periodic variation in the dielectric constant in the direction of
propagation.
Fig. 4 compares the approximate wave-vector curves obtained from (2) (solid lines) with the rigorous results taken
from the study in [14] (dashed lines) for the following: 1) a
symmetric slab waveguide surrounded by air and 2) with a silica
buffer layer and an air overlay. The slab thickness is 0.4 times
the period, and the air fraction is 0.3. The dielectric constant of
silica is 2.1. Fig. 4 shows that there is an excellent agreement
between these two results, particularly for the first few bands.
This approximate Bloch wave-vector diagram approach forms
the basis in explaining the close match between the measurement and calculation described in the next section.
III. F ABRICATION AND M EASUREMENTS
The 1-D PhC microcavities were etched into the ridge
waveguides that are fabricated from an SOI wafer with a

SHI et al.: MULTIWAVELENGTH TRANSMISSION MICROCAVITY IN SOI PLANAR RIDGE WAVEGUIDES

2209

TABLE I
QUALITY FACTORS AND RELATIVE TRANSMITTIVITY AT RESONANCE OF SIMPLE 1-D PhC MICROCAVITIES
ETCHED INTO 2.4-m-WIDE SOI RIDGE WAVEGUIDES

350-nm-thick silicon layer. A 2.4-m-wide ridge waveguide


was defined in the Si upper layer by conventional photolithography and subsequent reactive ion etching with a mixed gas
of trifluoromethane (CHF3 ) and sulfur hexafluoride (SF6 ). The
ridge height was estimated to be 300 nm, i.e., almost the full
thickness of the Si layer. While such waveguides are multimoded, the field profile of the propagating modes in the vertical
direction is strongly confined and well approximated by that
of the fundamental guided mode of the equivalent air/Si/silica
slab. This complies with the conditions in applying (2) to the
microcavity design.
The microcavity pattern was written into a 200-nm-thick
polymethyl methacrylate (PMMA) layer by electron-beam
lithography and transferred into the waveguide ridge by further
reactive ion etching. In order to maintain the accuracy of the
PhC structure dimensions during device fabrication and to prevent metal contamination of the RIE reactor, the PMMA layer
was used directly in protecting the unpatterned area without
using the ICP power during the RIE process. The depth of the
resulting air gaps was 200 nm, limited by the 1 : 1 etch
rates of Si and PMMA for the processing conditions used.
This complies with the condition of strong dielectric contrast
in the 1-D PhC in order to apply (2). Finally, the sample
was thinned by mechanical polishing to 150 m from the
backside and cleaved to produce optical quality facets to the
ridge waveguides.
Since the PMMA etch mask may result in erosion of the
edges of the Si blocks in the PhCs during RIE, the impact of
such structural degradation on microcavity performance was
investigated by first measuring the resonant behavior of a
number of simple waveguide microcavities fabricated by the
same process. The simple microcavities comprised of a half
wavelength (design wavelength = 1550 nm) Si central cavity
layer surrounded by Si/air Bragg reflectors without adiabatic
tapers intended to match the Bloch modes with the cavity
modes and the guided modes of the access waveguides [8], [9].
The 1-D PhC Bragg reflectors comprised of rectilinear Si
blocks that are separated by air gaps running across the whole
width of the waveguide ridge. The optical lengths of the Si
blocks and air gaps were one-quarter wavelength.

The device transmittance was measured by stepping a


C-band tunable diode laser source (TLS) in the spectral range of
the resonant peak and by measuring the transmitted light with
an optical power meter. By using this method, measurements
of Q and relative peak transmittivity Trel are limited by the
step resolution of the tunable laser: 1 pm in the case of the
instrument used in this paper.
Table I summarizes the measurements of Q and the peak
relative transmittivity Trel obtained by the laser wavelength
stepping method for simple microcavities with 2.5, 3.5, or
4.5 period quarter-wave Si/air Bragg reflectors. The Q-values
are compared very favorably with the measurements reported
in the literature for other waveguide-based 1-D PhC microcavities [5][8], with only well-optimized structures having
systematically higher Q [9], [16]. Although radiation recycling
may contribute to the measured Q [17], such high values also
indicate that there is only one dominant cavity mode that is most
likely derived from the fundamental mode of the nominally
multimoded 2.4-m-wide access ridge waveguides.
The measured Q and Trel of the microcavities with
4.5 period Bragg reflectors are lower than for the structures with
3.5 period mirrors, indicating that, while some degradation of
device performance from the scattering loss occurred, the fabrication methods were suitable in producing high-performance
microcavities.
Next, the wave-vector method was used in the design of a
waveguide-based microcavity with three resonant wavelengths
determined by the complex periodicity of the 1-D PhC. The
device structure is shown schematically in Fig. 1(c) and comprised of 1-D PhC reflectors of period A surrounding another
1-D PhC of period B (B > A ). The introduction of the
inner 1-D PhC results in a nonsymmetric structure with longer
air gaps in the central cavity region and the number of periods in
the Bragg reflectors differing. Based on the findings in Table I,
the numbers of periods in the Bragg reflectors were limited to
2 and 2.5 in order to maximize the transmittivity. The period
of the inner 1-D PhC and the air fraction were both optimized to produce transmission peaks in the telecommunications
bands at 1310- and 1550-nm wavelength with nearly equal
transmittivity.

2210

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 25, NO. 8, AUGUST 2007

Fig. 5. Scanning electron micrograph of a fabricated multichanneltransmission planar ridge waveguide filter based on an SOI wafer.

A scanning electron microphotograph of the fabricated structure is shown in Fig. 5. From left to right, the structure consists
of a two-period PhC reflector of measured period A = 493 nm
(design value of A = 500 nm) and a microcavity comprising
a two-period PhC of measured period B = 985 nm (design
value of B = 1000 nm). This is followed by a second reflector formed by a two and half-period PhC of period A .
The measured air fractions of periods A and B are 0.67 and
0.74, respectively, against design values of 0.72 and 0.73. The
discrepancy between the design and measured feature sizes
arises primarily from the limit on the exposure resolution of
the electron beam lithography process.
Light from a supercontinuum source consisting of a piece
of endless single-mode PhC fiber [18] was coupled into the
ridge waveguide through a microlensed taper. The transmitted light then was measured using an optical power meter.
Fig. 6(a) shows the transmission spectrum measured by an optical spectral analyzer (OSA). Three transmission peaks occur
at wavelengths of 1168, 1321, and 1562 nm. The positions of
the latter two peaks agree with the 1310- and 1550-nm design
specification of the microcavity to better than 1%.
Fig. 6(b) shows the calculated wave-vector curves of a
1-D PhC slab with a period of B = 985 nm. The air fraction
is 0.74. Values of the resonant wave vectors kB found from
(1), with Lp = 0, and the measured transmission frequencies
f = A / taken from Fig. 6(a) are indicated on the wavevector curves. The three measured transmission wavelengths
correspond to the fourth, fifth, and sixth resonant modes of the
cavity. Good agreement between the measured and calculated
results based on (2) is achieved with errors less than 3.5%.
The low spectral power density of the supercontinuum source
and correspondingly large resolution bandwidth of the OSA
needed for adequate detector sensitivity prevented accurate
measurements of the Q of both the resonance peaks at 1321 and
1562 nm. However, the C-band TLS enabled accurate measurements of the longer wavelength resonance. Fig. 7 compares the
relative transmittivity of the resonance at 1562-nm wavelength,
which is measured using the TLS with a wavelength step
resolution of 0.01 nm, with the data from Fig. 6(a) for the
corresponding spectral range. The Q and Trel measured by
the TLS are 230 and 0.11, respectively, indicating reasonable
control of the loss, despite the width of the air gaps in the inner
1-D PhC and their nonoptimum depth for minimizing scattering
[14]. Since the full width at half maximum of the resonance
peaks at 1321 and 1562 nm measured by the supercontinuum
source/OSA combination is comparable, at 10 nm, it can
be inferred that the Q and Trel of the resonance at 1321-nm
wavelength are broadly similar to those of the 1562-nm peak.

Fig. 6. (a) Transmission spectrum of the fabricated multichannel-transmission


planar ridge waveguide filter. (b) Comparison of the measured and calculated
resonant frequencies in the wave-vector diagram. The fourth, fifth, and sixth
resonant modes are labeled. The wave vector kB comes from (2) with f =
A /, A = 510 nm, B = 1012 nm, and B = 0.75.

Fig. 7. Comparison of the relative transmission in the spectral range of the


resonance at 1562-nm wavelength measured by the C-band tunable diode laser
(heavy line) and by the supercontinuum source and optical spectrum analyzer
(light line). The spectra have been offset to allow straightforward comparison
of the line shape.

This validates the design process based on the simple model


described above. While the Q-values for the resonant peaks
at 1321- and 1562-nm wavelength are low compared with the

SHI et al.: MULTIWAVELENGTH TRANSMISSION MICROCAVITY IN SOI PLANAR RIDGE WAVEGUIDES

2211

TABLE II
COMPARISON OF THE NORMALIZED RESONANT WAVELENGTHS

state of the art [9], [16], ultrahigh Q is not a requirement of the


communications application envisaged since a broad passband
is desirable to prevent data corruption.
IV. C ONCLUSION
A novel multiwavelength waveguide-based 1-D PhC microcavity has been realized using a simple design procedure based
on a wave-vector diagram approach.
Table II summarizes the results of the work described above.
First, the normalized resonant wavelengths A / obtained
from the simplified approach described in Section II to multilayer stack (details in the caption of Fig. 2) are compared with
the results of the rigorous TMM. The discrepancies between the
results of the two techniques are shown in brackets. Second, the
normalized resonant wavelengths of a planar slab waveguide
obtained from the new model are compared with those obtained from the measurement [Fig. 6(a)]. Again, the difference
between the results of the new model and measurements are
shown in brackets.
Excellent agreement is obtained in both cases, demonstrating
the validity of applying the simple wave-vector approach to the
design of complex 1-D PhC microcavities etched into ridge
waveguides with high index contrast. In the case of the complex multitransmission microcavity described here, the period
and air fraction of the 1-D PhCs have been optimized for
simultaneous transmission of data in the 1310- and 1550-nm
wavelength optical communications bands with nearly equal
transmittivity.
ACKNOWLEDGMENT
The authors would like to thank K. Hingerl of the University
of Linz, Linz, Austria, for collaboration and support.

[6] T. F. Krauss, B. Vgele, C. R. Stanley, and R. M. De La Rue, Waveguide


microcavity based on photonic microstructures, IEEE Photon. Technol.
Lett., vol. 9, no. 2, pp. 176178, Feb. 1997.
[7] D. J. Ripin, K.-Y. Lim, G. S. Petrich, P. R. Villeneuve, S. Fan,
E. R. Thoen, J. D. Joannopoulos, E. P. Ippen, and L. A. Kolodziejski,
One-dimensional photonic bandgap microcavities for strong optical
confinement in GaAs and GaAs/Alx Oy semiconductor waveguides,
J. Lightw. Technol., vol. 17, no. 11, pp. 21522160, Nov. 1999.
[8] D. Peyrade, E. Silberstein, P. Lalanne, A. Talneau, and Y. Chen, Short
Bragg mirrors with adiabatic modal conversion, Appl. Phys. Lett.,
vol. 81, no. 5, pp. 829831, Jul. 2002.
[9] P. Velha, J. C. Rodier, P. Lalanne, J. P. Hugonin, D. Peyrade, E. Picard,
T. Charvolin, and E. Hadji, Ultra-compact silicon-on-insulator ridge
waveguide mirrors with high reflectance, Appl. Phys. Lett., vol. 89,
no. 17, pp. 171 121171 123, Oct. 2006.
[10] K. Srinivasan, P. E. Barclay, and O. Painter, Fabrication tolerant high
quality factor photonic crystal microcavities, Opt. Express, vol. 12, no. 7,
pp. 14581463, Apr. 2004.
[11] Y. Akahane, T. Asano, B.-S. Song, and S. Noda, Fine tuned high Q
photonic crystal nanocavity, Opt. Express, vol. 13, no. 4, pp. 12021214,
Feb. 2005.
[12] H. Y. Lee, S. J. Cho, and G. Y. Nam, Multiple-wavelength-transmission
filters based on Si-SiO2 one-dimensional photonic crystals, J. Appl.
Phys., vol. 97, no. 10, p. 103 111, May 2005.
[13] P. S. J. Russell and T. A. Birks, Bloch wave optics in photonic
crystals: Physics and applications, in Photonic Band Gap Materials,
C. M. Soukoulis, Ed. Norwell, MA: Kluwer, 1996, pp. 7191.
[14] D. Gerace and L. C. Andreani, Gap maps and intrinsic diffraction
losses in one-dimensional photonic crystal slabs, Phys. Rev. E, Stat.
Phys. Plasmas Fluids Relat. Interdiscip. Top., vol. 69, no. 5, p. 056 603,
May 2004.
[15] J. A. McCaulley, V. M. Donnelly, M. Vernon, and I. Taha, Temperature
dependence of the near-infrared refractive index of silicon, gallium arsenide, and indium phosphide, Phys. Rev. B, Condens. Matter, vol. 49,
no. 11, pp. 74087417, Mar. 1994.
[16] M. W. Pruessner, T. H. Stievater, and W. S. Rabinovich, Integrated
waveguide Fabry-Perot microcavities with silicon/air Bragg mirrors,
Opt. Lett., vol. 32, no. 5, pp. 533535, Mar. 2007.
[17] P. Lalanne and J. P. Hugonin, Bloch mode engineering for high Q
small V microcavities, IEEE J. Quantum Electron., vol. 39, no. 11,
pp. 14301438, Nov. 2003.
[18] W. J. Wadsworth, N. Joly, J. C. Knight, T. A. Birks, F. Biancalana, and
P. S. J. Russell, Supercontinuum and four-wave mixing with Q-switched
pulses in endlessly single-mode photonic crystal fibres, Opt. Express,
vol. 12, no. 2, pp. 299309, Jan. 2004.

R EFERENCES
[1] E. Yablonovitch, Inhibited spontaneous emission in solid-state physics
and electronics, Phys. Rev. Lett., vol. 58, no. 20, pp. 20592062,
May 1987.
[2] J. D. Joannopoulos, R. Meade, and J. Winn, Photonic Crystals.
Princeton, NJ: Princeton Univ. Press, 1995.
[3] P. Russell, Photonic crystal fibers, Science, vol. 299, no. 5605, pp. 358
362, Jan. 2003.
[4] P. S. J. Russell, Designing photonic crystals, in Electron and Photon
Connement in Semiconductor Nanostructures, I. L. Nuovo Cimento, Ed.
Amsterdam, The Netherlands: IOS, 2003, pp. 79103.
[5] J. S. Foresi, P. R. Villeneuve, J. Ferrera, E. R. Thoen, G. Steinmeyer,
S. Fan, J. D. Joannopoulos, L. C. Kimerling, H. I. Smith, and E. P.
Ippen, Photonic-bandgap microcavities in optical waveguides, Nature,
vol. 390, no. 6656, pp. 143145, Nov. 1997.

Xiaohua Shi received the B.Eng. degree in electrical engineering from Xian
Jiaotong University, Xian, China, in 2003. Since 2003, he has been working
toward the Ph.D. degree with the Department of Electronic and Electrical
Engineering, University of Bath, Bath, U.K.
His research involves the design, nanofabrication, and characterization of
photonic crystal structure devices in silicon-on-insulator structures. He is currently with the University of Glasgow, Glasgow, U.K., researching microsystem
fabrication.

2212

Wei Ding received the B.Sc. degree in physics and the M.E. degree in electronics from Peking University, Beijing, China, in 1999 and 2002, respectively.
Since 2003, he has been working toward the Ph.D. degree with the Centre for
Photonic and Photonic Materials, Department of Physics, University of Bath,
Bath, U.K.
His current researches involve microfabrications on tapered optical fibers and
applications of plasmonics in scanning near-field optical microscopes.

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 25, NO. 8, AUGUST 2007

Duncan W. E. Allsopp (M89) received the B.Sc. degree in physics and the
M.Sc. and Ph.D. degrees from the University of Sheffield, Sheffield, U.K.,
in 1971, 1974, and 1977, respectively.
From 1977 to 1979, he was with Ferranti Electronics, Ltd., developing highspeed Si bipolar transistors. From 1979 to 1984, he was with the University of
Manchester Institute of Science and Technology, Manchester, U.K., researching
defects in semiconductors, and from 1984 to 1986, he was with British
Telecom Research Laboratories, Martlesham Heath, U.K. In 1986, he joined
the University of York, York, U.K., where he established a group researching
photonic devices. Since 1999, he has been in the Optoelectronics Group,
University of Bath, Bath, U.K., where he is currently a Royal Academy of
Engineering/Leverhulme Trust Senior Research Fellow, where he continues his
research into photonics.

Você também pode gostar