Conformally Back-Filled, Non-close-packed Inverse-Opal Photonic Crystals**

Similar documents
ZnS-based photonic crystal phosphors fabricated using atomic layer deposition

Sacrificial-Layer Atomic Layer Deposition for Fabrication of Non- Close-Packed Inverse-Opal Photonic Crystals**

Atomic layer deposition in porous structures: 3D photonic crystals

ATOMIC LAYER DEPOSITION FOR PHOTONIC CRYSTAL DEVICES

Luminescent and Tunable 3D Photonic Crystal Structures

Properties of Inverse Opal Photonic Crystals Grown By Atomic Layer Deposition

Supplementary Information

Seminar: Structural characterization of photonic crystals based on synthetic and natural opals. Olga Kavtreva. July 19, 2005

Fabrication of Silicon Inverse Woodpile Photonic Crystals**

7-2E. Photonic crystals

Low-cost, deterministic quasi-periodic photonic structures for light trapping in thin film silicon solar cells

Red luminescence from Si quantum dots embedded in SiO x films grown with controlled stoichiometry

Synthesis of inverse opals

Cristaux 3D. Fabriquer de tels objets aux longueur d'ondes optiques???

Solar Cells and Photosensors.

Mater. Res. Soc. Symp. Proc. Vol Materials Research Society

Enlargement of Grains of Silica Colloidal Crystals by Centrifugation in an

Large-Size Liftable Inverted-Nanobowl Sheets as Reusable Masks for Nanolithiography

OPTIMIZATION OF ALD GROWN TITANIA THIN FILMS FOR THE INFILTRATION OF SILICA PHOTONIC CRYSTALS. A Dissertation Presented to The Academic Faculty

Dispersion characteristics of silicon nanorod based carpet cloaks

Modification of Opal Photonic Crystals Using Al 2 O 3 Atomic Layer Deposition

Modified spontaneous emission from erbium-doped photonic layer-by-layer crystals

High Pressure Chemical Vapor Deposition to make Multimaterial Optical Fibers

Electric Field Distribution of Nanohole Thin Gold Film for Plasmonic Biosensor: Finite Element Method

Assembly of Highly Ordered Three-Dimensional Porous Structure with Nanocrystalline TiO 2 Semiconductors

Optical properties of a three-dimensional silicon square spiral photonic crystal

Fabrication of photonic band-gap crystals

Fabrication of annular photonic crystals by atomic layer deposition and sacrificial etching

Polycrystalline Silicon Produced by Joule-Heating Induced Crystallization

Supplementary Figure S1. Scheme for the fabrication of Au nanohole array pattern and

WHEN THE idea of photonic bandgap (PBG) materials

Metodi di fabbricazione PhC

Growth and Doping of SiC-Thin Films on Low-Stress, Amorphous Si 3 N 4 /Si Substrates for Robust Microelectromechanical Systems Applications

Supporting Information

Preparation and characterization of Co BaTiO 3 nano-composite films by the pulsed laser deposition

INVESTIGATION ON STRUCTURAL, OPTICAL, MORPHOLOGICAL AND ELECTRICAL PROPERTIES OF LEAD SULPHIDE (PbS) THIN FILMS

Growth Of TiO 2 Films By RF Magnetron Sputtering Studies On The Structural And Optical Properties

Supporting Information

Development of Amorphous Silicon Solar Cells with Plasmonic Light Scattering

Spectral emission imaging to map photonic properties below the crystal surface of 3D photonic crystals

Core-shell diamond-like silicon photonic crystals from 3D polymer templates created by holographic lithography

RightCopyright 2006 American Vacuum Soci

Self-Assembled 3d Photonic Crystals For Applications In Optical Communications

Crystallographic Characterization of GaN Nanowires by Raman Spectral Image Mapping

CHAPTER 4. SYNTHESIS OF ALUMINIUM SELENIDE (Al 2 Se 3 ) NANO PARTICLES, DEPOSITION AND CHARACTERIZATION

Enhanced Light Trapping in Periodic Aluminum Nanorod Arrays as Cavity Resonator

Supplementary Figure 1 TEM of external salt byproducts. TEM image of some salt byproducts precipitated out separately from the Si network, with

Heteroepitaxy of Monolayer MoS 2 and WS 2

Growth and formation of inverse GaP and InP opals

Silver Diffusion Bonding and Layer Transfer of Lithium Niobate to Silicon

Macroporous silicon: Homogeneity investigations and fabrication tolerances of a simple cubic three-dimensional photonic crystal

Synthesis and Characterization of DC Magnetron Sputtered ZnO Thin Films Under High Working Pressures

Versatile multi-layered metal-oxide inverse opal fabrication for photocatalytic applications

Trench Structure Improvement of Thermo-Optic Waveguides

1. Photonic crystal band-edge lasers

Rare Earth Doping of Silicon-Rich Silicon Oxide for Silicon-Based Optoelectronic Applications

Artificially inscribed defects in opal photonic crystals

Slow DNA Transport through Nanopores in Hafnium Oxide Membranes

Fundamentals of X-ray diffraction and scattering

Title: Localized surface plasmon resonance of metal nanodot and nanowire arrays studied by far-field and near-field optical microscopy

Electronic structure and x-ray-absorption near-edge structure of amorphous Zr-oxide and Hf-oxide thin films: A first-principles study

Study of The Structural and Optical Properties of Titanium dioxide Thin Films Prepared by RF Magnetron sputtering

Excimer Laser Annealing of Hydrogen Modulation Doped a-si Film

Reviewers' Comments: Reviewer #1 (Remarks to the Author)

for New Energy Materials and Devices; Beijing National Laboratory for Condense Matter Physics,

Facet-Selective Epitaxy of Compound Semiconductors on

Relation between Microstructure and 2DEG Properties of AlGaN/GaN Structures

Applications of Successive Ionic Layer Adsorption and Reaction (SILAR) Technique for CZTS Thin Film Solar Cells

Modeling Of A Diffraction Grating Coupled Waveguide Based Biosensor For Microfluidic Applications Yixuan Wu* 1, Mark L. Adams 1 1

Supplementary Information: Hybrid polymer photonic crystal fiber with integrated chalcogenide. glass nanofilms

لبا ب ةعماج / ةيساسلأا ةيبرتلا ةيلك ة لجم

ALD of Scandium Oxide from Tris(N,N -diisopropylacetamidinato)scandium and Water

PHOTOLUMINESCENCE AND SURFACE MORPHOLOGY OF NANOSTRUCTURED POROUS SILICON

This journal is The Royal Society of Chemistry S 1

Thermal Annealing Effects on the Thermoelectric and Optical Properties of SiO 2 /SiO 2 +Au Multilayer Thin Films

Nanofabrication with Laser Holographic Lithography for Nanophotonic Structures

ALD TiO 2 coated Silicon Nanowires for Lithium Ion Battery Anodes with enhanced Cycling Stability and Coulombic Efficiency

Simple fabrication of highly ordered AAO nanotubes

Defense Technical Information Center Compilation Part Notice

Development of photonic and thermodynamic crystals conforming to sustainability conscious materials tectonics

Confocal Microscopy of Electronic Devices. James Saczuk. Consumer Optical Electronics EE594 02/22/2000

Department of Chemistry, University of California, Davis, California 95616, USA 2

Fs- Using Ultrafast Lasers to Add New Functionality to Glass

ON THE PHOTOMAGNETIC EFFECT IN CdTe THIN FILMS EVAPORATED ONTO UNHEATED SUBSTRATES

Supporting Information

Effect of negative rf bias on electrophotographic properties of hard diamond-like carbon films deposited on organic photoconductors

Polycrystalline and microcrystalline silicon

SUPPLEMENTARY INFORMATION

Deposited by Sputtering of Sn and SnO 2

Optical Properties of CNT Arrays Growth in Porous Anodic Alumina Templates

Supporting Information. Solution-Processed 2D PbS Nanoplates with Residual Cu 2 S. Exhibiting Low Resistivity and High Infrared Responsivity

PbS NANO THIN FILM PHOTOCONDUCTIVE DETECTOR

Synthesis and Characterization of Zinc Iron Sulphide (ZnFeS) Of Varying Zinc Ion Concentration

TEM Study of the Morphology Of GaN/SiC (0001) Grown at Various Temperatures by MBE

Supporting Information. for Efficient Low-Frequency Microwave Absorption

Tunable Nanoscale Plasmon Antenna for Localization and Enhancement of Optical Energy. Douglas Howe

Tunable photonic bandgap structures for optical interconnects

Microscopic Structural Analysis of Advanced Anode Material for Lithium Battery

Compositional Profile of Graded VCSEL DBRs

Structure color. Photonic Structures: Discovery, Replication & Application. Yunyun Dai

Transcription:

DOI: 10.1002/adma.200501077 Conformally Back-Filled, Non-close-packed Inverse-Opal Photonic Crystals** By Jeffrey S. King, Davy P. Gaillot, Elton Graugnard, and Christopher J. Summers* Photonic crystals (PCs) offer far greater control over the generation and propagation of light than any other material structures. [1,2] Their defining characteristics, omnidirectional and directional photonic bandgaps (PBGs), can be manipulated to enable effects such as low losses in optical circuits and control of spontaneous emission. PCs are also creating much excitement because of their ability to form flat bands, yielding phenomena such as slow light and negative refraction. [3,4] The infiltration and inversion of synthetic-opal templates has been established as a promising method for obtaining the periodic structure and refractive-index contrast required in PCs. [5 9] In addition, topological tuning by shifting the distribution and filling fraction of a high-dielectric material within PCs offers a way to significantly change their photonic-band properties. For example, increased tunability, functionality, and bandgap properties can be accomplished through advanced architectures such as multilayered inverse opals and non-close-packed (NCP) inverse opals. [10 13] NCP geometries differ significantly from inverse opals by the formation of extended air cylinders between neighboring air spheres, as shown in Figure 1a, whereas, in the more-limited inverse opal, the air spheres are connected only by narrow sinter necks. As reported by Doosje et al., silicon NCP architectures are predicted to yield a 100 % increase in the width of the omnidirectional PBG between the eighth and ninth photonic bands. [12] In this paper, we show that modification of PC-template topology by both heat treatment and multiple conformal infiltrations facilitates precise control and optimization of photonic-band properties. Static tuning is possible by precisely controlled backfilling of inverse opals to increase the filling fraction of dielectric material. [14] However, inverse opals have inherently limited filling-fraction tunability because the narrow sinter necks quickly close with backfilling. To solve this limitation, we report the implementation of a new, two-step atomic layer deposition (ALD) infiltration process to form a [*] Prof. C. J. Summers, Dr. J. S. King, [+] D. P. Gaillot, Dr. E. Graugnard School of Materials Science and Engineering Georgia Institute of Technology 771 Ferst Drive, Atlanta, GA 30332-0245 (USA) E-mail: chris.summers@mse.gatech.edu [+] Present address: Department of Chemical Engineering, Stanford University, 381 North South Mall, Stanford, CA 94305-5025, USA. [**] This work was supported by the U.S. Army Research Office contract DAAD19-01-1-0603. Figure 1. a) Dielectric distribution in i) close-packed and ii) NCP inverse opals. Note cylindrical connections (air cylinders) between air spheres labeled in (ii), and sinter necks labeled in (i). b) NCP inverse-opal fabrication route: i) heavily sintered opal template, ii) resulting inverse opal, and iii) NCP inverse opal after atomic layer deposition backfilling. TiO 2 NCP inverse opal. Significantly, this new process allowed a static tunability of 400 nm in the position of the directional bandgap. In the first step, the void space available for infiltration was reduced to 25 % of the original volume by controllably collapsing the opal template by sintering followed by ALD infiltration, a method that is capable of high-finesse deposition of dense films within a nanoporous template. This facilitated the formation of an ultralow-filling-fraction inverse opal (5.8 %) with 193 nm diameter sinter necks. In the second step, backfilling of the low-volume-fraction inverse opal resulted in large tuning of the photonic-band properties and, ultimately, the formation of an NCP structure. Because of its many applications in biosensing, solar cells, catalysis, and environmental cleanup, optically active and tunable TiO 2 structures are of high interest. In this study, silica opal templates were first grown on silicon substrates in a manner similar to that of Park and coworkers, as described elsewhere. [15,16] The steps that were next taken to form an NCP inverse opal are illustrated in Figure 1b. First, a 10 lm thick opal template with 460 nm sphere diameter was sintered at 1000 C for 3 h (Fig. 1b, step i), which is considerably longer and at a higher temperature than the normal sintering condition of 800 C for 2 h, in order to partially collapse the template and increase the sinter-neck diameter. TiO 2 was then conformally deposited in the very small air voids (< 10 nm) using ALD, a technique that has re- Adv. Mater. 2006, 18, 1063 1067 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1063

cently shown high potential for PC fabrication. [11,16 20] The ability of ALD to fully and homogeneously infiltrate nanoporous structures was critical to achieving uniform infiltration of the narrow air channels (approximately four times smaller than conventional opals) of the sintered opal. The infiltrated sample was next heated to 400 C and held for 2 h to convert the as-deposited amorphous TiO 2 to the anatase phase. The top surface was then removed using an ion mill to expose the silica template, and etching in a 2 % HF solution resulted in the complete removal of the spheres. This first stage resulted in the formation of an ultralow filling fraction ( 5.8 vol.-%), structurally stable inverse opal, with a 220 % greater sinter neck diameter than possible in a typical inverse opal (Fig. 1b, step ii). In the second stage of the experiment, the TiO 2 PC s structure and optical properties were precisely tuned by backfilling the modified inverse-opal template with increasing TiO 2 depositions. The conformal nature of ALD allowed the periodicity of the inverse opal to be maintained, even for high levels of backfilling. Furthermore, the large sinter-neck diameter permitted significant backfilling, resulting in the formation of cylindrical (or hyperboloid-like) connections between the air spheres, thus yielding a well-formed NCP inverse opal, as shown in Figure 1b, step iii. This conformal model differs from that of Figure 1a, panel ii, a geometrical model, which is the same as the dielectric function used for band calculations by Doosje et al. [12] and Fenollosa and Meseguer. [13] Figure 2 shows cross-sectional scanning electron microscopy (SEM) images through the (111) plane of the NCP inverse opals. A schematic of the predicted 2D geometry of the air regions for each step is superimposed on the left of the images. The inverse opal produced by infiltration of a heavily sintered 460 nm opal followed by HF etching is shown in Figure 2a. As expected, the overall lattice constant was reduced, and measurement of the periodicity of the structure in the image revealed a 400 nm spacing between repeat units in the (111) plane. The resulting air-sphere diameter was 415 nm, slightly larger than the periodicity, but less than the original sphere size of 460 nm. Most significantly, the image indicates a 193 nm average diameter of the sinter necks, indicating formation of large pathways for gas flow within the structure. This structure is superior to a normal inverse opal because it increased by 220 % the amount of backfilling possible before the pores closed, improving the static tunability of the PC. After backfilling with 240 ALD cycles, the structure in Figure 2b was formed, in which the air spheres and sinter necks have been significantly reduced in diameter, and hyperboloid-like inter-sphere connections have formed. The airsphere and hyperboloid (sinter neck) radii decreased further after backfilling with an additional 320 ALD cycles (Figs. 2c and d). The formation of hyperboloid necks, air-sphere size tuning, and change in TiO 2 filling fraction, while maintaining a constant lattice constant, dramatically shows the degree of PC structural tuning achievable with this technique. To probe the resulting photonic-band structures in the C L direction, specular reflectivity at 15 incidence to the (111) planes was measured during each processing step of the NCP Figure 2. SEM images of ion-milled cross sections of the (111) plane of a) a TiO 2 inverse opal, formed from a heavily sintered 460 nm SiO 2 opal, b,c) an NCP inverse opal formed after 240 (b) and 560 (c) TiO 2 ALD backfilling cycles, and d) a higher-magnification image of the structure in (c). In (a), the dark shaded region indicates air-sphere overlap, as illustrated by the presence of large sinter necks. In (b,c,d), dark shaded regions highlight formation of hyperboloid-shaped connections between the air spheres of the new structure. 1064 www.advmat.de 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2006, 18, 1063 1067

showed that, by careful heavy sintering, not only the pore volume could be reduced by a factor of four, resulting in a 3D network of 10 nm pores, but also that conformal ALD can be used to achieve high-finesse template geometries. Remarkably, even with such a low filling fraction, the inverse opal was structurally stable, indicating a high TiO 2 density. This inverse-opal template was next infiltrated with amorphous TiO 2 in steps of 40 ALD cycles. Figure 3b shows the extremely wide range of tuning achievable after backfilling with 240 and 640 ALD cycles. A shift in the directional bandgap of 370 nm was realized, all from the same sample. Comparison of the evolution of the reflectivity spectra during the continual tuning of this NCP structure with the positions of the calculated (111) direction (C L) fundamental bandgaps (Bragg peak) is shown in Figure 4a. Figure 3. Reflectivity spectra for a) the first stage of fabrication: i) a heavily sintered SiO 2 opal, ii) a heavily sintered SiO 2 opal infiltrated with 340 cycles ( 17 nm) of TiO 2, and iii) an inverse NCP opal with 400 nm (111) periodicity and 5.8 % TiO 2 filling fraction. b) For the second stage of fabrication: i) a 5.8 % filling fraction inverse NCP opal, and after infiltration with b) 240 ALD cycles ( 12 nm) of TiO 2 and c) 640 ALD cycles ( 33 nm). (Specular reflectivity with a 15 incidence was measured). inverse opal. Figure 3a, curve i shows the spectrum of the heavily sintered opal, with a sharp, symmetrical Bragg reflectivity peak at 885 nm, and a sharp, lower-intensity peak at 440 nm attributed to flat band effects. As discussed by Míquez et al. [21] and Galisteo-López and López, [22] face-centered-cubic opals (and their derivatives) show reflectivity peaks not only in the presence of a bandgap, but also when the bands flatten. At these points, the group velocity slows, resulting in an increase in the refractive index and a resultant increase in reflectivity. After complete infiltration (Fig. 3a, curve ii), the peak remained sharp and symmetrical and shifted to 955 nm, which is much less than typically observed for infiltration of a conventional opal due to the reduced available air void volume. The flat band peak shifted to 475 nm and increased in intensity, close to that of the Bragg peak. After removing the SiO 2 spheres (Fig. 3a, curve iii), the Bragg peak remained sharp and well defined, and shifted to 739 nm, indicating a TiO 2 filling fraction of 5.8 % (calculated from the Bragg optical diffraction equation). Two higher-order peaks were also present at 392 and 427 nm. This data Figure 4. a) Comparison of peak positions (dark squares) and widths (open circles) (Dk/k) measured from reflectivity spectra with calculated positions of the second and third photonic bands (dotted lines) for the (111) (C L) direction, and b) measured relative peak widths and TiO 2 volume fractions for a TiO 2 NCP inverse opal as a function of ALD backfilling. Figure 4a shows the position of the calculated C L directional bandgaps (dotted lines) and the measured Bragg-peak positions (solid circles) and widths (hollow circles). [23] Both the peak positions and widths matched the predicted bandgap positions very well. The Bragg-peak Dk/k ratios were calculated from the measured reflectivity spectra after each infil- Adv. Mater. 2006, 18, 1063 1067 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advmat.de 1065

tration step by division of the full width at half maximum (FWHM) of the peak by its wavelength. The Dk/k ratio is a reliable indicator of changes in the PBG width. Secondary contributions to peak width, such as disorder and sphere polydispersity, are not expected to impact the tuning shown in this work since the highly conformal nature of the ALD process does not introduce additional disorder. Thus, the intrinsic peak broadening will remain constant after the sintering step. The relative peak widths (Dk/k) as a function of ALD cycles and the corresponding filling fractions are shown in Figure 4b. As shown in the bottom curve, the filling fraction increased linearly with ALD cycles until the last 60 80 cycles. In this region, a decrease in the slope of the filling-fraction curve indicated that the cylinders had closed, and further infiltration was not possible. The top curve shows that the peak width clearly displayed a maximum of 12.7 % at approximately 280 320 ALD cycles ( 20 % TiO 2 filling fraction), after which it decreased slightly to 11.6 %. However, after 440 ALD cycles (ca. 27 % TiO 2 filling fraction), the peak width increased with successive cycles, reaching a value of 14 % for the sample backfilled with 580 TiO 2 ALD cycles (36 % filling fraction). Additional infiltrations up to 640 ALD cycles ( 36 nm) and a 39 % filling fraction yielded a subsequent decrease in peak width and diminished gains in the TiO 2 filling fraction. In summary, a two-step ALD process was used to form nonclose-packed inverse opals by 1) conformally infiltrating heavily sintered silica opals with TiO 2 and etching the spheres with hydrofluoric acid, and 2) conformally backfilling the resulting inverse opal. Extensive sintering of the silica-opal template increased the sinter-neck diameter by 220 % due to sphere coalescence, with a consequent fourfold decrease in the available internal pore size and volume. These results demonstrate that ALD, a surface-limited, highly controllable growth technique, has the unique ability to fully and uniformly infiltrate porous structures, even within extended nanoscale channels (< 10 nm), which is critical to successful infiltration and inversion of the template. The inverse opal that resulted from infiltration of this template had a very low TiO 2 filling fraction (5.8 %), was structurally stable, and more significantly, had approximately 193 nm diameter sinter-neck regions connecting the air spheres. This topological transformation was essential for achieving high levels of backfilling (39 %) of the inverse opal and successful formation of hyperboloid-like connections between the air spheres that comprise the inverse opal. A 70 % increase in the TiO 2 filling fraction was attained over that possible without heavy pre-sintering (23 %). [17] The large range of TiO 2 volume fractions confirmed the wide variety of structures possible and the high degree of Bragg-peak tuning, which is nearly 400 nm wider than the entire visible spectral region, and larger than has ever been reported. These observations dramatically demonstrate the degree of structural changes achievable by this technique. This study is the first to demonstrate the power of a two-stage ALD process to uniquely enable the precise formation and tuning of complex PC structures. We expect this technique to be applicable to a broad range of PCs and to enable precise control and optimization of parameters that are critical for tuning the photonic-bandgap properties and dispersion characteristics. Experimental Monodisperse silica colloids were formed using the Stöber method and dispersed in ultrapure deionized (DI) water [24]. Silica opals were next grown by forced sedimentation in a confinement cell [15]. After sintering the opal films in air at 1000 C for 3 h, TiO 2 was grown within the opal in a custom-built flow-style hot-wall ALD reactor. Conformal growth was accomplished using alternating 8 s pulses of TiCl 4 and H 2 O, each separated by a 20 s, 225 sccm N 2 purge, while maintaining the substrate at 100 C. The vacuum level within the reactor was 500 mtorr during pulses and 200 mtorr during purges (1 Torr 133 Pa). The precursor gases, derived from room-temperature liquid sources, were introduced to a N 2 carrier gas using computer-controlled solenoid valves. Post-deposition heat treatment was performed at 400 C in the ALD reactor with a constant N 2 flow of 225 sccm. The surface of the heat-treated infiltrated opal was next ion-milled for several minutes with a 4 kev Ar + beam at 15 incidence. To remove the silica spheres, the milled sample was immersed in a 2 % HF solution for 1 h. The surface was next rinsed with DI water and dried in an oven at 110 C. Backfilling of the inverse opal utilized the same ALD infiltration routine. All SEM images were acquired using a LEO 1530 scanning electron microscope. Reflectivity measurements were performed using a Beckman DU640 spectrophotometer. Photonic-band diagrams were calculated using the finite-difference time-domain method. For the calculated band diagrams in Figure 4, an average refractiveindex value (n) for visible wavelengths was used (n anatase =2.65, n amorphous = 2.45 at 500 nm) [25,26]. When calculating the photonicband structures reported in this paper, a conformal model was used for generation of the required dielectric functions, which exactly simulates the ALD growth topology. This produced a dielectric function that matches the real structures, taking into account the actual shapes of the air spheres and the connecting regions between them that result after conformal backfilling. Received: May 25, 2005 Final version: October 24, 2005 Published online: March 22, 2006 [1] S. John, Phys. Rev. Lett. 1987, 58, 2486. [2] E. Yablonovitch, Phys. Rev. Lett. 1987, 58, 2059. [3] K. Sakoda, Optical Properties of Photonic Crystals, Springer, New York 2001. [4] C. M. Soukoulis, Photonic Crystals and Light Localization in the 21st Century, Kluwer Academic, Dordrecht, The Netherlands 2001. [5] J. E. G. J. Wijnhoven, W. L. Vos, Science 1998, 281, 802. [6] H. M. Yates, W. R. Flavell, M. E. Pemble, N. P. Johnson, S. G. Romanov, C. M. Sotomayor-Torres, J. Cryst. Growth 1997, 170, 611. [7] S. G. Romanov, N. P. Johnson, A. V. Fokin, V. Y. Butko, H. M. Yates, M. E. Pemble, C. M. S. Torres, Appl. Phys. Lett. 1997, 70, 2091. [8] A. Blanco, E. Chomski, S. Grabtchak, M. Ibisate, S. John, S. W. Leonard, C. López, F. Meseguer, H. Míguez, J. P. Mondla, G. A. Ozin, O. Toader, H. M. van Driel, Nature 2000, 405, 437. [9] Y. A. Vlasov, X.-Z. Bo, J. C. Sturm, D. J. Norris, Nature 2001, 414, 289. [10] F. García-Santamaría, M. Ibisate, I. Rodríguez, F. Meseguer, C. López, Adv. Mater. 2003, 15, 788. [11] J. S. King, D. Heineman, E. Graugnard, C. J. Summers, Appl. Surf. Sci. 2005, 244, 511. 1066 www.advmat.de 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2006, 18, 1063 1067

[12] M. Doosje, B. J. Hoenders, J. Knoester, J. Opt. Soc. Am. B 2000, 17, 600. [13] R. Fenollosa, F. Meseguer, Adv. Mater. 2003, 15, 1282. [14] H. Míguez, N. Tétreault, S. M. Yang, V. Kitaev, G. A. Ozin, Adv. Mater. 2003, 15, 597. [15] S. H. Park, D. Qin, Y. Xia, Adv. Mater. 1998, 10, 1028. [16] J. S. King, C. W. Neff, C. J. Summers, W. Park, S. Blomquist, E. Forsythe, D. Morton, Appl. Phys. Lett. 2003, 83, 2566. [17] J. S. King, E. Graugnard, C. J. Summers, Adv. Mater. 2005, 17, 1010. [18] A. Rugge, J. S. Becker, R. G. Gordon, S. H. Tolbert, Nano Lett. 2003, 3, 1293. [19] J. S. King, C. W. Neff, S. Blomquist, E. Forsythe, D. Morton, C. J. Summers, Phys. Status Solidi B 2004, 241, 763. [20] M. Scharrer, X. Wu, A. Yamilov, H. Cao, R. P. H. Chang, Appl. Phys. Lett. 2005, 86, 1. [21] H. Míguez, V. Kitaev, G. A. Ozin, Appl. Phys. Lett. 2004, 84, 1239. [22] J. F. Galisteo-López, C. López, Phys. Rev. B: Condens. Matter Mater. Phys. 2004, 70, 035 108. [23] For comparison with the calculations, the measured peak positions and widths were converted to values expected for normal incidence by using the Bragg equation for optical diffraction. [24] W. Stöber, A. Fink, E. Bohn, J. Colloid Interface Sci. 1968, 26, 62. [25] Handbook of Optical Constants of Solids (Ed: E. D. Palik), Vols. 1 3, Academic, San Diego, CA 1998. [26] D. Heineman, M. S. Thesis, Georgia Institute of Technology 2004. Adv. Mater. 2006, 18, 1063 1067 2006 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.advmat.de 1067