Fabrication-tolerant high quality factor photonic crystal microcavities

Similar documents
Air-holes radius change effects and structure transitions in the linear photonic crystal nanocavities

Photonic Crystal Nanocavities for Efficient Light Confinement and Emission

Polarization control and sensing with two-dimensional coupled photonic crystal microcavity arrays. Hatice Altug * and Jelena Vučković

Appendix. Photonic crystal lasers: future integrated devices

Investigation on Mode Splitting and Degeneracy in the L3 Photonic Crystal Nanocavity via Unsymmetrical Displacement of Air-Holes

Tailoring of the resonant mode properties of optical nanocavities in two-dimensional photonic crystal slab waveguides

Two-dimensional porous silicon photonic crystal light emitters

Spectral properties of photonic crystal double heterostructure resonant cavities

Photonic crystal with multiple-hole defect for sensor applications

THE spontaneous emission coupling factor ( factor) of

Photonic devices for quantum information processing:

General recipe for designing photonic crystal cavities

Angular and polarization properties of a photonic crystal slab mirror

Defect-based Photonic Crystal Cavity for Silicon Laser

The influence of material absorption on the quality factor of photonic crystal cavities


Single Semiconductor Nanostructures for Quantum Photonics Applications: A solid-state cavity-qed system with semiconductor quantum dots

Confinement of band-edge modes in a photonic crystal slab

Quantum Cascade Photonic Crystal Surface-Emitting Injection Lasers

Genetic optimization of photonic bandgap structures

Design of quantum cascade microcavity lasers based on Q factor versus etching depth

Design of a high-q air-slot cavity based on a width-modulated line-defect in a photonic crystal slab

Behavior of light at photonic crystal interfaces

Design of mid-ir and THz quantum cascade laser cavities with complete TM photonic bandgap

SUPPLEMENTARY INFORMATION

Continuous room-temperature operation of optically pumped InGaAs/InGaAsP microdisk lasers

Far- and Near-Field Investigations on the Lasing Modes in Two-Dimensional Photonic Crystal Slab Lasers

Photonic band structure of ZnO photonic crystal slab laser

Band structure of honeycomb photonic crystal slabs

Nanomaterials and their Optical Applications

Research on the Wide-angle and Broadband 2D Photonic Crystal Polarization Splitter

Dielectric-Band Photonic Crystal Nanobeam Lasers

High-Q Defect-Free 2D Photonic Crystal Cavity from Random Localised Disorder

FINITE-DIFFERENCE FREQUENCY-DOMAIN ANALYSIS OF NOVEL PHOTONIC

Optimum Access Waveguide Width for 1xN Multimode. Interference Couplers on Silicon Nanomembrane

Room temperature continuous wave operation and controlled spontaneous emission in ultrasmall photonic crystal nanolaser

Polarization control of defect modes in threedimensional woodpile photonic crystals

OPTI510R: Photonics. Khanh Kieu College of Optical Sciences, University of Arizona Meinel building R.626

Nanocomposite photonic crystal devices

Modified spontaneous emission from a twodimensional photonic bandgap crystal slab

Strong Coupling between On Chip Notched Ring Resonator and Nanoparticle

Title. Author(s)Nagasaki, Akira; Saitoh, Kunimasa; Koshiba, Masanori. CitationOptics Express, 19(4): Issue Date Doc URL.

Enhancement mechanisms for optical forces in integrated optics

Surface plasmon waveguides

A photonic crystal superlattice based on triangular lattice

Room temperature continuous wave lasing in InAs quantum-dot microdisks with air cladding

arxiv:quant-ph/ v3 20 Apr 2005

Photonic crystal waveguide-mode orthogonality conditions and computation of intrinsic waveguide losses

A direct analysis of photonic nanostructures

Spontaneous emission rate of an electric dipole in a general microcavity

Splitting of microcavity degenerate modes in rotating photonic crystals the miniature optical gyroscopes

A COMPACT POLARIZATION BEAM SPLITTER BASED ON A MULTIMODE PHOTONIC CRYSTAL WAVEGUIDE WITH AN INTERNAL PHOTONIC CRYSTAL SECTION

Defect modes of a two-dimensional photonic crystal in an optically thin dielectric slab

Chapter 5. Effects of Photonic Crystal Band Gap on Rotation and Deformation of Hollow Te Rods in Triangular Lattice

Photonic Crystals: Periodic Surprises in Electromagnetism. You can leave home without them. Complete Band Gaps: Steven G.

Resonantly Trapped Bound State in the Continuum Laser Abstract

Fabrication and optical measurements of silicon on insulator photonic nanostructures

Photonic crystals. Semi-conductor crystals for light. The smallest dielectric lossless structures to control whereto and how fast light flows

Advanced techniques Local probes, SNOM

Photonic crystal fiber with a hybrid honeycomb cladding

Design of a Multi-Mode Interference Crossing Structure for Three Periodic Dielectric Waveguides

A new method for sensitivity analysis of photonic crystal devices

RECENT innovations in nanoscale optical technologies

Study of Propagating Modes and Reflectivity in Bragg Filters with AlxGa1-xN/GaN Material Composition

Directional emitter and beam splitter based on self-collimation effect

SUPPLEMENTARY INFORMATION

Enhancing the Rate of Spontaneous Emission in Active Core-Shell Nanowire Resonators

Stimulated Raman amplification and lasing in silicon photonic band gap nanocavities

Emission pattern control and polarized light emission through patterned graded-refractiveindex coatings on GaInN light-emitting diodes

Mode coupling and cavity quantum-dot interactions in a fiber-coupled microdisk cavity

Out-of-plane scattering from vertically asymmetric photonic crystal slab waveguides with in-plane disorder

Strongly coupled single quantum dot in a photonic crystal waveguide cavity

Emission Spectra of the typical DH laser

Multi-cycle THz pulse generation in poled lithium niobate crystals

Guided and defect modes in periodic dielectric waveguides

Circularly polarized thermal emission from chiral metasurface in the absence of magnetic field

Single-photon nonlinearity of a semiconductor quantum dot in a cavity

Photonic crystal laser threshold analysis using 3D FDTD with a material gain model

Study on Quantum Dot Lasers and their advantages

Photonic Micro and Nanoresonators

Photonic/Plasmonic Structures from Metallic Nanoparticles in a Glass Matrix

arxiv: v1 [quant-ph] 20 Feb 2008

Tailorable stimulated Brillouin scattering in nanoscale silicon waveguides.

1. Fabrication. Lukáš Ondič a, Marian Varga a, Karel Hruška a, Jan Fait a,b and Peter Kapusta c

Quantum Dot Lasers Using High-Q Microdisk Cavities

Cavity QED: Quantum Control with Single Atoms and Single Photons. Scott Parkins 17 April 2008

Introduction to optical waveguide modes

SUPPLEMENTAL MATERIAL I: SEM IMAGE OF PHOTONIC CRYSTAL RESONATOR

HIGH-Q PHOTONIC CRYSTAL NANOBEAM CAVITY BASED ON A SILICON NITRIDE MEMBRANE INCOR- PORATING FABRICATION IMPERFECTIONS AND A LOW-INDEX MATERIAL LAYER

IN conventional optical fibers, light confinement is achieved

Finite-difference time-domain calculation of spontaneous emission lifetime in a microcavity

Nonlinear optical spectroscopy in one-dimensional photonic crystals. Abstract

Light Interaction with Small Structures

SUPPLEMENTARY INFORMATION

Quantum Dot Lasers. Jose Mayen ECE 355

Morphology-dependent resonance induced by two-photon excitation in a micro-sphere trapped by a femtosecond pulsed laser

Free-Space MEMS Tunable Optical Filter in (110) Silicon

Theory of Photonic Crystal Slabs by the Guided-Mode Expansion Method

Negative curvature fibers

Transcription:

Fabrication-tolerant high quality factor photonic crystal microcavities Kartik Srinivasan, Paul E. Barclay, and Oskar Painter Department of Applied Physics, California Institute of Technology, Pasadena, CA 91125, USA. phone: (626) 395-6269, fax: (626) 795-7258, e-mail: kartik@caltech.edu Abstract: A two-dimensional photonic crystal microcavity design supporting a wavelength-scale volume resonant mode with a calculated quality factor (Q) insensitive to deviations in the cavity geometry at the level of Q 2 10 4 is presented. The robustness of the cavity design is confirmed by optical fiberbased measurements of passive cavities fabricated in silicon. For microcavities operating in the λ = 1500 nm wavelength band, quality factors between 1.3-4.0 10 4 are measured for significant variations in cavity geometry and for resonant mode normalized frequencies shifted by as much as 10% of the nominal value. c 2004 Optical Society of America OCIS codes: (230.5750) Resonators; (140.5960) Semiconductor lasers References and links 1. O. Painter, R. K. Lee, A. Yariv, A. Scherer, J. D. O Brien, P. D. Dapkus, and I. Kim, Two-Dimensional Photonic Band-Gap Defect Mode Laser, Science 284, 1819 1824 (1999). 2. T. Yoshie, J. Vučković, A. Scherer, H. Chen, and D. Deppe, High quality two-dimensional photonic crystal slab cavities, Appl. Phys. Lett. 79, 4289 4291 (2001). 3. Optical Processes in Microcavities, R. K. Chang and A. J. Campillo, eds., (World Scientific, Singapore, 1996). 4. H. J. Kimble, Strong Interactions of Single Atoms and Photons in Cavity QED, Physica Scripta T76, 127 137 (1998). 5. P. Michler, A. Kiraz, C. Becher, W. Schoenfeld, P. Petroff, L. Zhang, E. Hu, and A. Imomoglu, A Quantum Dot Single-Photon Turnstile Device, Science 290, 2282 2285 (2000). 6. C. Santori, M. Pelton, G. Solomon, Y. Dale, and Y. Yamamoto, Triggered Single Photons from a Quantum Dot, Phys. Rev. Lett. 86, 1502 1505 (2001). 7. J. Vučković, M. Lončar, H. Mabuchi, and A. Scherer, Design of photonic crystal microcavities for cavity QED, Phys. Rev. E 65 (2002). 8. K. Srinivasan and O. Painter, Momentum space design of high-q photonic crystal optical cavities, Opt. Express 10, 670 684 (2002), http://www.opticsexpress.org/abstract.cfm?uri=opex-10-15-670. 9. H.-Y. Ryu, M. Notomi, and Y.-H. Lee, High-quality-factor and small-mode-volume hexapole modes in photoniccrystal-slab nanocavities, Appl. Phys. Lett. 83, 4294 4296 (2003). 10. K. Srinivasan, P. E. Barclay, O. Painter, J. Chen, A. Y. Cho, and C. Gmachl, Experimental demonstration of a high quality factor photonic crystal microcavity, Appl. Phys. Lett. 83, 1915 1917 (2003). 11. Y. Akahane, T. Asano, B.-S. Song, and S. Noda, High-Q photonic nanocavity in a two-dimensional photonic crystal, Nature 425, 944 947 (2003). 12. K. Srinivasan, P. E. Barclay, M. Borselli, and O. Painter, Optical fiber based measurement of an ultrasmall volume, high-q photonic crystal microcavity, submitted to Phys. Rev. Lett., Sept. 2003 (available at http://arxiv.org/quant-ph/abs/0309190). 13. O. Painter, K. Srinivasan, and P. Barclay, A Wannier-like Equation for Photon States of Locally Perturbed Photonic Crystals, Phys. Rev. B 68, 035214 (2003). 14. K. Srinivasan and O. Painter, Fourier space design of high-q cavities in standard and compressed hexagonal lattice photonic crystals, Opt. Express 11, 579 593 (2003), http://www.opticsexpress.org/abstract.cfm?uri=opex-11-6-579. 15. J. Knight, G. Cheung, F. Jacques, and T. Birks, Phase-matched excitation of whispering-gallery-mode resonances by a fiber taper, Opt. Lett. 22, 1129 1131 (1997). 16. Y. Tanaka, T. Asano, Y. Akahane, B.-S. Song, and S. Noda, Theoretical investigation of a two-dimensional photonic crystal slab with truncated air holes, Appl. Phys. Lett. 82, 1661 1663 (2003). (C) 2004 OSA 5 April 2004 / Vol. 12, No. 7 / OPTICS EXPRESS 1458

(a) (b) y x 1 µm Fig. 1. (a) FDTD calculated magnetic field amplitude ( B ) in the center of the optically thin membrane for the fundamental A 0 2 mode. (b) Scanning electron microscope image of a fabricated Si PC microcavity with a graded defect design (PC-5 described below). 1. Introduction Two-dimensional photonic crystal (PC) slab waveguide microcavities[1, 2] offer the promise of simultaneously exhibiting a high quality factor (Q) and an ultra-small, wavelength-scale modal volume (V eff ). These two parameters, which physically represent long photon lifetimes and large per photon electric field strengths, respectively, are key to microcavity-enhanced processes in nonlinear optics, quantum optics, and laser physics[3, 4, 5, 6]. Recent progress on PC microcavities has included theoretical work on the design of PC microcavities with predicted Q factors from 10 4 to 10 6 [7, 8, 9], and experimental work demonstrating Q factors in excess of 10 4 in InP-based lasers[10] and silicon membranes[11, 12]. A range of microcavity designs have been employed in these studies, and in many cases, the experimental achievement of high-q is predicated on the ability to fabricate the design with a small margin for error. For example, in Ref. [2], the discrepancy between the fabricated device and the intended design led to a theoretical degradation of Q from 3.0 10 4 to 4.4 10 3, consistent with the measured Q of 2.8 10 3. Extraordinary control over fabricated geometries has been demonstrated in recent work[11], where a shift of 60 nm in the positions of holes surrounding the cavity defect region reduced Qs as high as 4.5 10 4 by over an order of magnitude. Here, we discuss work on a PC microcavity[10, 12] that exhibits a degree of robustness, both theoretically and experimentally, to deviations from the nominal design sufficient for Qs above 10 4 to be maintained. This robustness in Q to changes in the PC cavity geometry is of practical importance for future experiments in the aforementioned disciplines, to provide insensitivity to fabrication imperfections, as well as to maintain the flexibility in cavity design required to form resonant modes with a prescribed field pattern and polarization. Radiative losses in planar waveguide two-dimensional PC defect microcavities can be separated into in-plane and out-of-plane components, quantified by the quality factors Q and Q, respectively, with the total radiative Q given by Q 1 = Q 1 + Q 1. Q is determined by the size and angular extent (in-plane) of the photonic bandgap, while Q is determined by the presence of in-plane momentum components (k) within the waveguide cladding light cone, which are not confined by total internal reflection at the core-cladding interface. In Ref. [8], PC microcavities were designed using two mechanisms to avoid radiative loss: (i) use of a mode that is odd about mirror planes normal to its dominant Fourier components, in order to eliminate the DC (k = 0) part of the in-plane spatial frequency spectrum and hence reduce vertical radiation loss, and (ii) use of a grade in the hole radius to further confine the mode and reduce in-plane (C) 2004 OSA 5 April 2004 / Vol. 12, No. 7 / OPTICS EXPRESS 1459

(a) 0.35 0.31 (b) r/a (c) 0.27 0.23 0.36 A B C D E A,B (d) r/a 0.32 0.28 1 4 2 56 0.24 3 7-5 -4-3 -2-1 0 1 2 3 4 5 x-axis position (a) -4-3 -2-1 0 1 2 3 4 y-axis position (a) Fig. 2. Grade in the normalized hole radius (r/a) along the central ˆx and ŷ axes of square lattice PC cavities such as those shown in Fig. 1. Cavity r/a profiles for (a,b) FDTD cavity designs and (c,d) microfabricated Si cavities. radiative losses. The resulting PC microcavity design within the square lattice creates a TE-like (magnetic field predominatly along ẑ) donor-type defect mode (labeled A 0 2 )1 as shown in Fig. 1(a). FDTD simulations of this resonant mode predict a Q-factor of 10 5 and an effective modal volume of V eff 1.2(λ/n) 3. We now show how use of mechanisms (i) and (ii) above create a level of robustness in the cavity design. Use of an odd symmetry mode to suppress vertical radiation loss is, at a basic level, independent of changes in the size of the holes defining the defect cavity. This feature has been confirmed in simulations of simple defect cavity designs in square lattice photonic crystals[8], where Q did not degrade below 10 4, despite significant changes (as much as 40%) in the size of the (two) central defect holes. Perturbations that cause the cavity to be asymmetric create a mode which, though not strictly odd, will be a perturbation to an odd mode, and hence will still largely suppress DC Fourier components and exhibit high Q. However, for the square lattice photonic crystal structures considered here, perturbations to the central defect hole geometry can result in a degradation in Q, due in part to the lack of a complete in-plane bandgap within the square lattice. This lack of a complete bandgap requires the defect geometry to be tailored so as to eliminate the presence of Fourier components in directions where the lattice is no longer reflective. This tailoring was achieved in Ref. [8] by a grade in the hole radius moving from the center of the cavity outwards. The grade, shown in Fig. 1, serves to help eliminate couplings to in-plane radiation modes along the diagonal axes of the square lattice (the M-point of the reciprocal lattice) where the PC is no longer highly reflective, while simultaneously providing a means to keep the in-plane reflectivity high along the ŷ axis (the direction of the mode s dominant Fourier components). The use of a large number of holes to define the defect region ensures that no single hole is responsible for creating the potential well that confines the resonant mode, making the design less susceptible to fluctuations in the size of individual holes. Instead, the continuous change in the position of the conduction band-edge resulting from the grade in hole radius creates an approximately harmonic potential well[13]. This smooth change in the position of the band-edge creates a robust way to mode-match between the central portion of 1 This label refers to the mode s symmetry classification and to it being the lowest frequency mode in the bandgap. (C) 2004 OSA 5 April 2004 / Vol. 12, No. 7 / OPTICS EXPRESS 1460

Table 1. Theoretical (PC-A through PC-E) and experimental (PC-1 through PC-7) normalized frequency (a/λ o ) and quality factor (Q) values for the A 0 2 mode of cavities with profiles shown in Fig. 2. Cavity d/a a/λ 0 Q Q Q PC-A 0.750 0.245 1.1 10 5 4.7 10 5 9.0 10 4 PC-B 0.750 0.245 1.1 10 5 2.6 10 5 7.5 10 4 PC-C 0.750 0.247 1.0 10 5 3.7 10 5 8.0 10 4 PC-D 0.750 0.253 8.6 10 4 3.0 10 5 6.7 10 4 PC-E 0.750 0.266 6.2 10 4 6.5 10 5 5.6 10 4 PC-1 0.879 0.241 - - 1.6 10 4 PC-2 0.850 0.255 - - 1.8 10 4 PC-3 0.850 0.251 - - 1.7 10 4 PC-4 0.842 0.251 - - 2.4 10 4 PC-5 0.842 0.249 - - 2.5 10 4 PC-6 0.800 0.263 - - 4.0 10 4 PC-7 0.800 0.270 - - 1.3 10 4 the cavity (where the mode sits) and its exterior. In other work[11], softening of this transition is achieved by adjusting the position of two holes surrounding the central cavity region (which consists of three removed air holes in a hexagonal lattice). This method can achieve high-q,but as mode-matching is achieved by tailoring only two holes it is more sensitive to perturbations than the adiabatic transition created by a grade in the hole radius. Finally, we note that even though a relatively large number of holes are modified to create the graded lattice, V eff is still wavelength-scale, and remains between 0.8-1.4(λ/n) 3 in all of the devices considered in this work. In addition, the methods used here to achieve robustness in Q are general and can be applied to cavities in other PC lattices[14]. To highlight these ideas, 3D FDTD simulations of cavities with varying grades and average normalized hole radius ( r/a) were performed. Figure 2(a)-(b) shows the grade in r/a along the central ˆx and ŷ axes for several designs (PC-A through PC-E), and Table 1 lists the calculated resonant frequency, vertical, in-plane, and total Q factors. In all of these simulations, Q remains close to 10 5, with PC-E showing more significant degradation largely as a result of the increased modal frequency (creating a larger-sized cladding light cone). In addition, an inappropriate choice of grade along the ˆx-axis can lead to increased in-plane losses via coupling to M-point modes. Nevertheless, the loss in any of the simulated devices did not cause Q to be reduced below 2 10 4. To test the sensitivity of the design to perturbations experimentally, cavities were fabricated in a d=340 nm thick silicon membrane through a combination of electron beam lithography, inductively-coupled plasma reactive ion etching, and wet etching. Figure 2(c)-(d) shows the values of r/a along the central ˆx and ŷ axes for a number of fabricated devices (PC-1 through PC-7), as measured with a scanning electron microscope (SEM). Cavities are passively tested[12] using an optical fiber taper[15], which consists of a standard single mode optical fiber that has been heated and stretched to a minimum diameter of 1-2 µm. At such sizes, the evanescent field of the fiber mode extends into the surrounding air, providing a means by which the cavity modes can be sourced and out-coupled. The fiber taper is spliced to the output of a fiber-pigtailed scanning tunable laser (1565-1625 nm) with 1 pm resolution, and is mounted (Fig. 3(a)) above and parallel to an array of PC cavities (Fig. 3(b)). When it is brought into close (C) 2004 OSA 5 April 2004 / Vol. 12, No. 7 / OPTICS EXPRESS 1461

(a) (b) linewidth, γ (nm) (c) 0.10 0.09 0.08 Transmission PC chip 1.00 0.98 γ=0.070nm 0.96 1620.5 1620.75 1621 wavelength (nm) (d) 0.09 0.08 0.07 0.06 Transmission 1.000 0.996 γ=0.047 nm 25 µm 0.992 1618.5 1618.75 1619 wavelength (nm) 0.07 0 100 200 300 400 500 600 700 800 z=taper-pc gap (nm) 0.05 0.04 0 100 200 300 400 500 600 700 800 z=taper-pc gap (nm) Fig. 3. (a) Schematic illustrating the fiber taper probe measurement setup. (b) SEM image of an array of undercut PC cavities (white box indicates position of one device within the array). (c) Measured data (blue dots) and exponential fit (red curve) for linewidth vs. taper-pc gap of the A 0 2 mode in PC-5. (Inset) Taper transmission for this device when the taper-pc gap is 350 nm. (d) Same as (c) for PC-6 (here, the taper transmission in the inset is shown when z=650 nm). The transmission curves are normalized relative to transmission in the absence of the PC cavity. proximity ( 500 nm) to the sample surface, and the polarization of the fiber taper is adjusted (via polarization-controlling padddle wheels) to be TE-polarized relative to the slab, evanescent coupling between the taper and cavity modes occurs. Figure 3(c)-(d) shows measurements for devices PC-5 and PC-6, which have significantly different r/a profiles (Fig. 2(c)-(d)). The inset of Fig. 3(c) shows the normalized taper transmission as a function of wavelength when the taper is 350 nm above cavity PC-5. By measuring the dependence of cavity mode linewidth (γ) on the vertical taper-pc gap ( z) (Fig. 3(c)), an estimate of the true cold-cavity linewidth (γ 0 ) is given by the asymptotic value of γ reached when the taper is far from the cavity. For PC-5, γ 0 0.065 nm, corresponding to Q 2.5 10 4. Figure 3(d) shows the linewidth measurement for PC-6. For this device, γ 0 0.041 nm, corresponding to a Q 4.0 10 4. As described in Ref. [12], the strength of the taper-pc coupling as a function of taper position can be used to estimate the spatial localization of the cavity field; these measurements closely correspond with calculations and for PC-6 are consistent with an FDTD-predictedV eff 0.9(λ/n) 3. These PC microcavities thus simultaneously exhibit a high-q factor that is insensitive to perturbations, and an ultra-small V eff. Linewidth measurements for each of the cavities PC-1 through PC-7 are compiled in Table 1. The robustness of the Q to non-idealities in fabrication is clearly evident. Though all of the devices exhibit a general grade in r/a, the steepness of the grade and the average hole radius ( r/a) vary considerably without reducing Q below 1.3 10 4. These high-q values are exhibited despite the fact that many cavities are not symmetric (the odd boundary condition is thus only approximately maintained), and the frequency of the cavity resonance varies over a 10% range, between a/λ o = 0.243-0.270. The measured Q values in Table 1 are still lower than predicted from simulations. This discrepancy is likely due in part to slightly angled etched sidewalls that have been shown in (C) 2004 OSA 5 April 2004 / Vol. 12, No. 7 / OPTICS EXPRESS 1462

calculations to lead to radiative coupling to TM-like modes[16]. This non-ideality helps explain why PC-1, which is closest in r/a value to the desired design (PC-A), does not exhibit the highest Q experimentally. In particular, we have observed that the sidewall angle is poorer for smaller sized holes. On the other end of the spectrum, cavities with the largest hole sizes such as PC-7, which may have more vertical sidewalls, also begin to exhibit higher vertical radiation loss as a result of a larger modal frequency and cladding light cone. In addition, surface roughness is a potential source of loss; for PC-6, which exhibited the highest Q value, a chemical resist stripping process was used (rather than a plasma de-scum) and may have produced a cleaner, smoother surface. In summary, the robustness in Q to errors in the in-plane design of a PC microcavity consisting of a graded square lattice of air holes is discussed. This property is confirmed both by FDTD simulations of devices where the steepness of the grade and the average hole radius are varied without degrading Q below 2 10 4, and in measurements of microfabricated Si cavities that exhibit Q factors between 1.3-4.0 10 4 over a wide range of device parameters. For these high-q cavities, current limitations on the Q factor appear to stem principally from slightly angled sidewalls and etched surface roughness, as opposed to errors in the in-plane shape or size of holes. This work was partly supported by the Charles Lee Powell Foundation. The authors thank M. Borselli for his contributions in building the taper test setup. K.S. thanks the Hertz Foundation for its financial support. (C) 2004 OSA 5 April 2004 / Vol. 12, No. 7 / OPTICS EXPRESS 1463