Skip to main content

Quasicrystals - Research Log

Quasicrystals



3D Photonic Image-Structure Projection Technology.























This website is interesting: http://unifiedfractalfield.com/

Negative Refraction and Imaging Using 12-fold-Symmetry Quasicrystals

Zhifang Feng, Xiangdong Zhang,  Yiquan Wang, Zhi-Yuan Li, Bingying Cheng, and Dao-Zhong Zhang

Institute of Physics, Chinese Academy of Sciences, Beijing 100080, China Department of Physics, Beijing Normal University, Beijing 100875, China (Received 5 July 2004; published 23 June 2005)


Recently, negative refraction of electromagnetic waves in photonic crystals was demonstrated experi- mentally and subwavelength images were observed. However, these investigations all focused on the periodic structure. Here, we report a new theoretical and experimental finding that negative refraction can appear in some transparent quasicrystalline photonic structures. The photonic quasicrystals (PQCs) exhibit an effective refractive index close to ō°€1 in a certain frequency window. The index shows small spatial dispersion, consistent with the nearly homogeneous geometry of the quasicrystal. More interest- ingly, a superlens based on the 2D PQCs can form a non-near-field subwavelength image whose position varies with the source distance. These properties make PQCs promising for application in a range of optical devices. 


Looking For Possible Military applications:

Three-dimensional quasicrystalline photonic material with five-fold planar symmetry for visible and infrared wavelengths by holographic assembly of quasicrystalline photonic heterostructures


Department of Physics and Center for Soft Matter Research, New York University, New York, NY 10003, USA Current address: Department of Physics & Astronomy, The University of Texas at San Antonio, One UTSA Circle, San Antonio TX 78249, USA
*vkf002@my.utsa.edu

Abstract: In this paper, we investigate three-dimensional (3D) band gap properties of quasiperiodic structure. We successfully demonstrate the fabrication of a 3D dielectric quasicrystalline heterostructures with five-fold planar symmetry using the holographic optical tweezers technique. Light transmitted through this quasicrystal is collected using the spatially resolved optical spectroscopy technique for both visible and infrared wavelength bandwidths in a far-field region. We investigate and analyze the transmission spectra for the same wavelength bandwidths in a near-field region by using computer simulations. The computational modeling indicates that for both TE and TM modes of propagating light in the XY plane there is a clear transmission band-gap of around 50 nm wide centered at 650 nm. This indicates that there is a rotational symmetry in the constructed quasicrystal along its XY plane. Future directions and applications are discussed.


© 2013 Optical Society of America
OCIS codes: (060.3510) Lasers, fiber; (090.1760) Computer holography; (160.5298) Photonic crystals.
References and links
  1. S. E. Burkov, T. Timusk, and N. W. Ashcroft, “Optical conductivity of icosahedral quasi-crystals,” J. Phys. Condens. Matter 4(47), 9447–9458 (1992).
  2. J. D. Joannopoulos, R. D. Meade, and J. N. Winn, Photonic Crystals (Princeton University Press, 1995).
  3. M. Florescu, S. Torquato, and P. J. Steinhardt, “Complete band gaps in two-dimensional photonic quasicrystals,”
    Phys. Rev. B 80(15), 155112 (2009).
  4. L. Jia, I. Bita, and E. L. Thomas, “Level Set Photonic Quasicrystals with Phase Parameters,” Adv. Funct. Mater.
    22(6), 1150–1157 (2012).
  5. X. Zhang, Z. Q. Zhang, and C. T. Chan, “Absolute photonic band gaps in 12-fold symmetric photonic crystals,”
    Phys. Rev. B 63(8), 081105 (2001).
  6. M. C. Rechtsman, H.-C. Jeong, P. M. Chaikin, S. Torquato, and P. J. Steinhardt, “Optimized Structures for
    Photonic Quasicrystals,” Phys. Rev. Lett. 101(7), 073902 (2008).
  7. J. Xu, R. Ma, X. Wang, and W. Y. Tam, “Icosahedral quasicrystals for visible wavelengths by optical
    interference holography,” Opt. Express 15(7), 4287–4295 (2007).
  8. A. Ledermann, L. Cademartiri, M. Hermatschweiler, C. Toninelli, G. A. Ozin, D. S. Wiersma, M. Wegener, and
    G. von Freymann, “Three-dimensional silicon inverse photonic quasicrystals for infrared wavelengths,” Nat.
    Mater. 5(12), 942–945 (2006).
  9. Y. Roichman and D. G. Grier, “Holographic assembly of quasicrystalline photonic heterostructures,” Opt.
    Express 13(14), 5434–5439 (2005).
  1. E. R. Dufresne and D. G. Grier, “Optical tweezer arrays and optical substrates created with diffractive optical elements,” Rev. Sci. Instrum. 69(5), 1974–1977 (1998).
  2. D.G.Grier,“Arevolutioninopticalmanipulation,” Nature424(6950),810–816(2003).
  3. E.D.Palik,HandbookofOpticalConstantofSolids (Academic,1985).
  4. W. Man, Photonic Quasicrystals and Random Ellipsoid Packings: Experimental Geometry in Condensed Matter
    Physics (Princeton University PhD Dissertation, 2005).
  5. P.M.Chaikin,“PhotonicQuasicrystals, http://www.physics.nyu.edu/~pc86/.
  6. Version 4.3, “COMSOL Multiphysics Reference Guide,” http://www.comsol.com.
  7. W. Man, M. Megens, P. J. Steinhardt, and P. M. Chaikin, “Experimental measurement of the photonic properties
    of icosahedral quasicrystals,” 
  8. T.F.Krauss,R.M.D.L.Rue,andS.Brand,“Two-dimensionalphotonic-bandgapstructuresoperatingatnear
    infrared wavelengths,” Nature 383(6602), 699–702 (1996).
  9. J. D. Joannopoulos, P. R. Villeneuve, and S. Fan, “Photonic crystals: putting a new twist on light,” 
Introduction
One of the major challenges in photonic crystal technology lies in designing materials with specific absorption properties that are able to control and filter light over specific wavelength bandwidths. Such materials are useful for developing new types of photonic band gap filters. Among these materials, dielectric quasicrystalline heterostructures are particularly applicable for creating interesting diffraction effects for visible and infrared light due to their unique crystalline symmetries. Quasicrystals are ordered structures lacking of the translational periodicity of crystals. Interestingly, this deficiency allows quasicrystals to adopt a wide variety of unusual long-ranged rotational symmetries which are absent for crystals. In fact, the large number of effective lattice vectors in the reciprocal space provides quasicrystals with effective Brillouin zones [1]. This leads to the formation of photonic band gaps [2–4] for light traveling through dielectric quasicrystalline heterostructures [5,6], despite the relatively low ratio of dielectric constants among the constituent materials [7]. Recently, three- dimensional (3D) silicon inverse photonic quasicrystals for infrared wavelengths were designed by direct laser writing and subsequent silicon single-inversion approach [8]. Also, holographic assemblies of two-dimensional (2D) and 3D silica quasicrystalline heterostructures were demonstrated [9].
In this article, we present, a 3D assembly of five-fold symmetry silica quasicrystalline heterostructures using the holographic optical trapping technique [10,11]. The far-field forward scattered light through the constructed quasicrystal along Z-axis was collected by using spatially resolved optical spectroscopy technique. We also analyzed the near-field forward scattered light via computational modeling in order to reveal photonic band gap properties of the constructed quasicrystal. A three-dimensional finite element method (FEM) in commercially available COMSOL multiphysics software was used for the optical modeling of the proposed quasiperiodic structure. A plane electromagnetic wave was assumed to be incident normally along X, M (i.e. direction at a 45 degree angle with respect to X and Y axes), Y and Z-axes on the sample. The energy range of the incident radiation varied from 2.48 to 0.83 eV, which correspond to the wavelength range of ~500-1500 nm. Both the transverse electric (TE) and transverse magnetic (TM) modes were considered. In the notation used in this work, the TE mode is defined as the electric field of light perpendicular to a surface in which the incident light propagates, whereas the TM mode is defined as the magnetic field of light perpendicular to the same surface. Port boundary conditions were used in both the positive and negative Z, X, M and Y-directions and perfect electric and magnetic boundary conditions were used along different axis accordingly. We used a finer built-in free triangular mesh in COMSOL multiphysics. The minimum and the maximum element sizes of the triangular mesh were 4.42 nm and 0.103 Ī¼m respectively with the maximum element growth rate of 1.35, and 0.3 resolution of curvature. The relative tolerance was set to 0.001. The dielectric functions used for all the materials in this work can be found elsewhere.

OPTICAL MATERIALS EXPRESS
drive memory was used for simulations. Each simulation approximately required 24 hour time period for obtaining transmission spectrum for the energy range of interest.

2. Experimental details
2.1 Quasicrystal fabrication
We prepared polyacrylamide hydrogel from colloidal 1.5 Ī¼m diameter silica microspheres (Duke scientific Lot 5238), dispersed in an aqueous solution of Acrylamide, N, N′ - Methylenebisacrylamide and Diethoxyacetophenone (All Aldrich Electrophoresis grade), in the ratio of 180:12:1 by weight. A sample was prepared by dispensing 4 Ī¼L of the hydrogel on the microscope slide and covering it with 18x18 cover glass. The cover glass was glued to the microscope slide with 5 minute epoxy. Thus, we obtained the hydrogel sample of 9 Ī¼m in thickness which was sealed and ready for use. Next, the sample was mounted on the microscope’s stage in order to experimentally build colloidal quasicrystalline material and analyze its optical properties. All vertices of the quasicrystal were generated by specially designed computer algorithm [13]. Computer generated vertices were then projected as holograms through a high-numerical-aperture microscope lens, creating large 3D arrays of optical traps. Silica particles immersed in the hydrogel were arranged into quasicrystal (consisting of 190 spheres in 5 layers) by these 3D optical traps. This particular domain consists of 190 spheres in 5 layers. Details regarding the design and spacing between layers and vertices.

Quasicrystal constructed from spherical silica particles. (a) Colloidal silica particles trapped in three dimensional configurations with holographic optical traps (Media 1). Brightness of a particle corresponds to different crystalline layer in Z-axis. From (b) to (e): schematic representations of different projections of the quasicrystal. Colors represent different heights of the crystalline planes in Z-axis. Orange, black, red, green and blue correspond to 3.7, 2.9, 0.4, −0.4 and −3.7 Ī¼m respectively.
2.2 Experimental setup
The schematic of the experimental setup used for optical study is shown in Fig. 2. A frequency-doubled diode-pumped solid state laser (Coherent Verdi 5W) generated light at 532 nm which was imprinted with phase-only hologram with the help of a liquid crystal spatial light modulator (SLM) (Hamamatsu X7690-16). The imprinted laser beam was translated to the input pupil of a 100x NA 1.4 SPlan Apo oil immersion objective positioned in an inverted optical microscope (Nikon TE-2000U). The objective focused the modified laser beam into traps and revealed images of trapped objects.

OPTICAL MATERIALS EXPRESS
Fig. 2. The experimental setup consists of a Laser, SLM, Microscope, Charge Coupled Device (CCD) Camera, Optical Fiber and Spectrometer all connected to the Computer. Laser beam passes through a series of optical devices creating an array of a three dimensional traps in microscope’s conventional imaging plane. (a) Real-time spectrum of forward scattered light from quasicrystal is collected by the optical fiber mounted to the second eyepiece port and connected to the Spectrometer. (b) Real-time image of the constructed quasicrystal, obtained from the CCD camera mounted to the first eyepiece port of the microscope.
After trapping silica particles and arranging them into quasicristalline structure, the sample was illuminated by an ultraviolet light which rapidly photopolymerized the structure. Video demonstrating experimental steps and details during the fabrication process can be seen in Media 1. In order to study optical properties of the constructed quasicrystal, the sample was illuminated by the visible and infrared light from microscope’s condenser. The resulting forward scattered light from the quasicrystal was collected by 500 Ī¼m diameter optical fiber connected to USB400 Plug-and-Play Miniature Fiber Optic Spectrometer and NIR512, Near- infrared Spectrometer both by (Ocean Optics Inc.). Note that the fabricated 3D photonic structure may reveal interesting optical properties in different directions due to the presence of 2D (fivefold in XY plane) quasicrystalline symmetries with several plane layers in Z direction.

3. Results and discussion
3.1 Experimental Measurements
The transmission spectra for light propagating along the Z-axis through the quasicrystal measured for both visible and infrared wavelength bandwidths in the far-field region are shown in Fig. 3. We find that the first harmonic in the infrared spectrum is located at 1175 nm while the second harmonic in the visible spectrum is located at 585 nm. The first harmonic corresponds to ~857 nm in the medium which is comparable to the interplanar spacing between the layers of the quasicrystal as shown in Figs. 1(d) and 1(e) (orange-black, and red- green).


OPTICAL MATERIALS EXPRESS 1335

Visible and infrared transmission spectra of the quasicrystal sample obtained in far-field measurements along the Z-axis.
The maximum coefficients of the transmission spectra at visible and infrared wavelengths are 0.23 and 0.028, respectively. Large crystals with more layers can increase the interference pattern of the diffracted light and strengthen the intensity of the transmitted light. The construction of a crystal with large numbers of layers is restricted by the experimental setup because the amount of optical traps is limited by SLM. The transmission signal can be also improved by increasing the dielectric contrast of crystal-medium. In our case, the ratio of dielectric constants is ~1.10. We observe that the signal of the first harmonic in the infrared spectrum is weaker than the signal for the second harmonic located in the visible spectrum. This is caused by the optical lenses inside the microscope which are regular (Relay optics) rather than specially designed ones for infrared light.

3.2 Computer Simulations
We investigated light propagation in different directions in the proposed quasicrystalline structure using COMSOL multiphysics software. In order to reveal complete photonic band gaps of the quasicrystal; that is, a range of wavelengths over which light propagation is not permitted for all directions and polarizations, we investigate both TE and TM modes in different directions. Detailed information about differences between TE/TM modes and the band gap theory for one dimensional (1D), 2D and 3D photonic crystals and quasicrystals can be found in [14]. A band gap is defined as the transmission minima with the coefficient nearly equal to zero or less than the simulation tolerance (0.001). Figures 4(a) and 4(b) depict the calculated transmission spectra from the quasicrystal in the near field region for both the TE and TM modes of light propagating along X and Y axes in the XY plane. A clear band gap (around 100 nm wide) centered at 650 nm is observed. This indicates that there is rotational symmetry in XY plane of the crystal for both TE and TM polarizations.
Fig. 4. Transmission spectra calculated for both the TE and TM modes, for the light propagating along the (a) X-axis, (b) Y-axis, (c) Z-axis and (d) XY plane at 0, 45 and 90 degree angles corresponding to X, M and Y directions. The gap at 650 nm shows an evidence of the complete band gap for light propagating in XY plane. Arrows and dashed lines indicate the locations of band gaps.
#192295 - $15.00 USD Received 17 Jun 2013; revised 26 Jul 2013; accepted 30 Jul 2013; published 13 Aug 2013
(C) 2013 OSA 1 September 2013 | Vol. 3, No. 9 | DOI:10.1364/OME.3.001332 | OPTICAL MATERIALS EXPRESS 1336

Figure 4(c) shows the transmission spectra for light propagating in the Z direction with ZX(TE) and ZY(TM) polarizations. We observe clear band gaps at 1000 nm and 680 nm, respectively. However, there is no rotational symmetry for both of these directions and as a result gaps are shifted and appear at different wavelengths. Figure 4(d) corresponds to the transmission spectra for the propagation of light in XY-plane at three different angles. We observe a clear rotational symmetry band gap around 50 nm wide centered at 650 nm for both TE and TM polarization which clearly shows evidence of a complete band gap.
The required transmission spectrum is retrieved from calculated S-matrix using the port boundary conditions. The calculation details of the S-parameters can be found at [15].
We chose monodisperse silica spheres because of their lower optical absorption, higher density and commercial availability. Other materials with specific optical properties can also be used for holographic assembly of quasicrystalline heterostructures. These studies are first trials towards the complete 3D band gap investigations of dielectric quasicrystals for visible and infrared wavelength bandwidths, constructed with the holographic optical trapping technique. Further investigations require construction of larger quasicrystals with icosahedral symmetry and analysis of transmitted light at different angles [16].

The experimental techniques used in our studies have advantages compared to other experimental methods used for the fabrication of quasicrystals. For instance, W. Man et. al. [16] experimentally investigated band gap properties of icosahedral quasicrystal constructed by using stereolithography technique. The experimental techniques used in our studies can create better dielectric contrast between the constituent materials of a quasicrystal in different directions compared to stereolithography technique. We are trapping and fixing dielectric spherical particles at specific locations in the medium without connecting them. In the stereolithography fabrication method, vertices of constructed quasicrystal are connected with dielectric rods. This in return may reduce the dielectric contrast between the vertices, and affect the band gap properties of constructed 3D quasicrystal. Thus, using our experimental techniques, we can obtain more accurate band gap picture of 3D quasicrystals.
In general, controlling the propagation of light at visible and infrared wavelengths is an extremely important field of research due to recent advancements in photonic crystal technology [17, 18].
4. Conclusion
We demonstrated holographic assembly of 3D dielectric quasicrystalline heterostructures with five-fold planar symmetry. We collected far-field forward scattered light from the quasicrystal using the spatially resolved optical spectroscopy technique. The propagation of light in different directions for both TE and TM modes was studied using finite element method in COMSOL multiphysics software. We observed evidence of the complete band gap of the constructed quasicrystal in the near field region. Results from the experimental measurements and the computational modeling were obtained despite a small dielectric contrast between the constituent materials of the quasicrystal. In conclusion, the techniques used in this work are powerful tools for constructing 3D quasiperiodic structures and investigating their full band gap properties which could lead to the development of new types of 3D photonic band gap filters for industrial applications.
Acknowledgments
Zaven Ovanesyan greatly appreciates the inspiration and support provided by Prof. David Grier to carry out this research work at the Department of Physics at the New York University. Authors are grateful to Weining Man, Lucas Fernandez Seivane and Zurab Kereselidze for valuable assistance. This work was supported by the Brownian Transport Thru Modulated Potential Energy Landscapes Grant Number DMR-0451589 and FRG: Photonic Quasicrystals and Heterostructures Grant Number DMR-0606415.
#192295 - $15.00 USD Received 17 Jun 2013; revised 26 Jul 2013; accepted 30 Jul 2013; published 13 Aug 2013

Wolf Clan Media

I found these papers but cannot access most of them. If anyone with access (students, researchers, etc.) decides to help with the investigation. Please reach out. Thanks.

https://wolfclanentertainment.com/


http://www.sciencedirect.com/science/article/pii/S0921509300013058





$208 USD
https://books.google.com/books?id=6MqMjDPthG8C&lpg=PP2&ots=SLNjF_2VPY&dq=quasicrystals%20military%20applications&lr&pg=PP2#v=onepage&q&f=false

$60
https://books.google.com/books?id=fN0I6f1XyJgC&lpg=PA92&ots=EAsT700qoF&dq=recent%20trends%20in%20thermoelectric%20materials%20research%20SAIC&pg=PA92#v=onepage&q&f=false


Duke University PDF - Thermoelectric Materials

Possible applications - Vice news

NIST on Dan Schetman

NIST - Dec. 2001 
https://www.nist.gov/publications/nist-centennial-celebration-crystallographic-highlights


https://journals.aps.org/prl/abstract/10.1103/PhysRevLett.90.055501


Abstract

We propose a holographic design of five-beam symmetric umbrella configuration, where there are a central beam and four ambient beams symmetrically scattered around the central one with the same apex angle, for fabrication of three-dimensional photonic crystals with tetragonal or cubic symmetries, and systematically analyzed the band gap properties of resultant photonic crystals when the apex angle is continuously increased. Our calculations reveal that large complete photonic band gaps exist in a wide range of apex angle for a relatively low refractive index contrast. Specifically, the face-centered cubic structure with a relative band gap of 25.1% for Īµ = 11.9 can be obtained with this recording geometry conveniently where all the beams are incident from the same half-space. These results will provide us with more understanding of this important recording geometry and give guidelines to its use in experiments.
©2006 Optical Society of America

1. Introduction

Photonic crystals (PhCs) are structures in which the dielectric constant is periodically modulated on a length scale comparable to the desired wavelength of operation [12], and the resultant photonic dispersion may exhibit photonic band gaps (PBGs) which are useful in controlling light behavior. In the last decade much attention has been attracted to the fabrication of PhCs with complete PBGs. Various techniques such as the electron-beam lithography, self-assembly, multiphoton polymerization, and holographic lithography (HL) have been proposed and demonstrated with different levels of success [3–7]. Among them the method of HL has some unique features such as inexpensive volume recording.
In HL a desired geometrical structure is formed by multibeam interference with single or multiple exposures [8–11]. We have shown that all 14 three-dimensional (3D) Bravais lattices can be produced this way [12], and different beam designs will result in different PBG properties of the resultant structures [1314]. For example, the face-centered-cubic (fcc) lattice can be obtained by the interference of one central beam and three ambient beams symmetrically scattered around the former [1516], but the structure made in this geometry has only quite a narrow PBG [1617]. An alternative beam design was proposed to fabricate fcc lattice with a large complete PBG, but it requires four beams incident from two opposite surfaces of a sample [18–20], making it difficult to realize in practice. In addition, in real experiments the directions of interference beams may be slightly deviated from their theoretical values for many practical reasons [16]. Therefore a more extensive investigation into the effect of beam angle derivation on the resulting structures and then on the PBGs will be helpful for a clearer understanding of the structure formation with this geometry and for the control of experimental parameters. Recently, a research group fabricated some tetragonal and cubic PhCs with the use of phase masks and analyzed the PBGs of these structures; however, what was employed in their calculations is the woodpile model instead of the real holographic structure [21], and the two cases are usually not the same since in HL the actual structures including the cell’s shape and size and consequently the corresponding PBG are strongly dependent on the specific beam design and light intensity threshold selection.
In this paper we propose a five-beam symmetric umbrella configuration which is more convenient to make 3D PhCs with large complete PBGs. For instance, an fcc lattice with 25.1% relative band gap for a dielectric constant contrast 11.9:1 can be obtained with all the beams incident from the same side of the sample. Furthermore, a systematical study on the structure and band gap evolution for a continuously varying apex angle in this geometry is provided. These analyses and results may serve as guidelines in both theory and practice.

2. Recording geometry and resultant structures

The recording geometry with five-beam symmetric umbrella configuration we proposed here is shown in Fig. 1, where the central beam (C-beam) with wave vector K c is set along the z direction, while the four ambient beams (A-beams) of wave vectors K 1 to K 4 are in the plane yoz or xoz, respectively, with the same apex angle Īø with z axis. The five wave vectors can be expressed as functions of the apex angle,

K1=(2Ļ€Ī»)(sinĪø,0,cosĪø),K2=(2Ļ€Ī»)(0,sinĪø,cosĪø),

K3=(2Ļ€Ī»)(sinĪø,0,cosĪø),K4=(2Ļ€Ī»)(0,sinĪø,cosĪø),

Kc=(2Ļ€Ī»)(0,0,1).
This geometry can be realized conveniently with the use of a diffraction beam splitter (DBS) where a zero-order diffracted beam and four symmetric first-order diffracted beams are employed. For polarization, we choose the central beam to be circularly polarized and all the four A-beams linearly polarized. The corresponding unit polarization vectors are

e1=e3=(0,1,0),e2=e4=(1,0,0),ec=12(1,i,0).

Fig. 1. Symmetric umbrella recording geometry and the coordinate system used for calculations.
If we use the same amplitude EA for all the A-beams and the amplitude EC for C-beam, and adopt double-exposure approach with the C-beam, beams 1 and 2 open for the first exposure while the C-beam, beams 3 and 4 open for the second exposure, the two exposures have the same exposure time, the total intensity will be

I(r)=2EC2+4EA2+2ECEA{sin[2Ļ€Ī»(xsinĪø+(1cosĪø)z)]

+cos[2Ļ€Ī»(ysinĪø+(1cosĪø)z)]

+sin[2Ļ€Ī»(xsinĪø+(1cosĪø)z)]

+cos[2Ļ€Ī»(ysinĪø+(1cosĪø)z)]}.
To get the best contrast we should maximize the ratio of √2EC EA /(2EC2+4EA2), and this requirement leads to the optimized beam amplitude ratio EC /EA = √2. For brevity we may write the relative intensity of the spatially varying part in Eq. (3) as

Ī”I(r)=sin[2Ļ€Ī»(xsinĪø+(1cosĪø)z)]

+cos[2Ļ€Ī»(ysinĪø+(1cosĪø)z)]

+sin[2Ļ€Ī»(xsinĪø+(1cosĪø)z)]

+cos[2Ļ€Ī»(ysinĪø+(1cosĪø)z)].
A common approach in HL is the use of a negative photoresist material such as SU-8. In this case the “underexposed” region can be selectively removed using a developer substance which leaves the “overexposed” region intact. The developed photoresist is then infiltrated with SiO2and burned away, leaving a daughter inverse template. Finally the daughter template is inverted by high temperature infiltration with silicon [20]. In the following analysis we will consider this case. If we denote the light intensity threshold in Eq. (4) as I t and assume that the region of Ī”I > I t is filled with a material of high refractive index while the other region is air, we may obtain a periodic microstructure whose concrete shape is determined by the specific apex angle Īø and I t.
In general the lattice structures resulting from Eq. (3) or Eq. (4) are tetragonal symmetric structures. The continuous increase of apex angle Īø leads to continuous variation of primitive vectors, reciprocal vectors and the irreducible Brillouin zone of the resultant structure. Our calculations show that the Brillouin zone changes from a small tetragonal cake spreading out on the xy plane when Īø is very small to a long tetragonal pillar along the z axis when Īø is close to 180° (see Fig. 2). When the apex angle Īø reaches 70.53° (corresponding to c/a = √2, where a is the period of the interference pattern along x or y direction and c is that in z direction), the structure has fcc symmetry, similar to the diamond structure [22]. If the Īø is near 70.53°, a lattice with face-centered-tetragonal (fct) symmetry is obtained. Figure 3 (a) and 3(b) show the real fcc structure and its primitive cell constructed by five-beam symmetric umbrella configuration when Īø = 70.53°, which obviously differ from the rhombohedral structure with fcc symmetry and its primitive cell [Fig. 3(c) and 3(d)] formed by four-beam symmetric umbrella configuration [12] when the apex angle is 38.94°. If the apex angle achieves the value of Īø = 90° (corresponding to c/a = 1), a body-centered-cubic (bcc) lattice is obtained. When the value of Īø is near 90°, the structure has body-centered-tetragonal (bct) symmetry.

Fig. 2. Evolution of irreducible Brillouin zone (Movie, 155K). The irreducible Brillouin zones of tetragonal structures for Īø = 80° (a), Īø = 90° (b), and Īø = 100° (c).

Fig. 3. (a). The real fcc structure formed by five-beam symmetric umbrella configuration when Īø = 70.53° and I t = 1.39 in Eq. (4), the corresponding filling ration is 21.7%; (b) the primitive cell of the fcc structure shown in (a); (c) the rhombohedral structure with fcc symmetry fabricated by four-beam symmetric umbrella configuration when the apex angle is 38.94°; and (d) the primitive cell of the structure shown in (c).

3. Band gap calculations

The plane-wave expansion method [2324] has been used to study PBG properties of the structures of this kind and search for the corresponding optimum volume filling ratio f yielding maximum relative PBG for each apex angleĪø. Calculations reveal that full photonic band gaps exist over a very wide range of apex angle with a relatively low refractive index contrast needed to open them. Figure 4 represents the relative gap sizes of optimized structures for the apex angle range of 50° < Īø < 115° with a dielectric constant contrast of 11.9 to 1 corresponding to silicon in air. From Fig. 4 we can see clearly that there are complete PBGs above 10 % in the range of 52° < Īø < 112°, and even larger PBGs above 20% for 59° < Īø < 92°. The maximum relative gap size of 25.1% appears at Īø = 70.53° for the fcc structure, and the relative gap size for the bcc structure formed when Īø = 90° is 21.3%. As an example, the band structure of the fcc lattice is given in Fig. 5, obviously a large complete PBG, from 0.330Ļ‰a/2Ļ€c to 0.425Ļ‰a/2Ļ€c(where c is the light velocity), exists between the second and third bands. Comparing the band gap result with that of woodpile structure fabricated by woodpile model [21], one can find two results have similar trend that the band gap size is a function of c/a defined in last section or Īø, and the fcc structure has the biggest gap size.

Fig. 4. Relative band gap of optimized structures as a function of apex angle for 50°< Īø < 115° when Īµ = 11.9. The solid symbols are the date for fcc and bcc structures when Īø = 70.53° and Īø = 90°, respectively. When the value of Īø is near 70.53° and 90° the fct and bct symmetric structures can be obtained.

Fig. 5. Photonic band structure for the fcc structure with Īø = 70.53°. The position of the high symmetry points together with the irreducible Brillouin zone are shown in the inset.
The effect of the dielectric refractive index n of the structure formed by this five-beam symmetric umbrella configuration on the band gap size has also been investigated. As a special case, in Fig. 6, we illustrate the variation of relative band gaps with different filling ratios and different refractive index contrasts for the fcc symmetric structure formed when Īø = 70.53°. In this case the minimum refractive index required to open a complete photonic band gap is slightly less than 1.95, lower than 2.05 obtained in one work of band gap calculation for the woodpile model as the lowest requirement to open the gap [25]. When the refractive index is 3.6 and the filling ratio is 21.5%, the relative band gap reaches as high as 27.3%. In addition, it is worth noting that the range of filling ratio yielding a complete band gap for each given refractive index is fairly wide, especially for the high refractive index. For example, when the dielectric refractive index is 3.4 or 3.6, a complete relative band gap larger than 10% can be obtained over a wide range of filling ratio, from 10% to 53%.

Fig. 6. Variation of relative band gaps with different filling ratios for different refractive index contrasts for the fcc symmetric structure with the apex angle Īø = 70.53°.

4. Conclusions

In summary, we have designed a five-beam symmetric umbrella configuration for the fabrication of 3D photonic crystals using holographic lithography method. The photonic structures with different symmetries, including face-centered-tetragonal, face-centered-cubic, body-centered-tetragonal and body-centered-cubic, can be produced with this geometry of different apex angles. The theoretical analysis indicates that complete band gaps exist for the structures formed this way over a very wide range of apex angle. Particularly, the fcc and bcc symmetric structures can be obtained when Īø = 70.53° and 90°, respectively; and their corresponding relative band gap sizes are as high as 25.1% and 21.3%, respectively, for a dielectric constant contrast 11.9:1. When the value of Īø is near 70.53° and 90° the fct and bct symmetric structures can be obtained, respectively. Furthermore, the dielectric constant contrast or equally the refractive index contrast of the structures of this kind required to open complete PBGs is quite low (n ≥ 1.95), and the filling ratio range for a certain structure to assure complete PBGs is fairly large. These discussions give us more understanding of this configuration and its advantages in applications. For example, we can fabricate an fcc lattice with a large complete PBG with all the interference beams arranged in the same half-space of the sample, which is more convenient compared with the geometry where the beams are incident from two opposite sides; the requirement for an exact theoretical apex angle may be relaxed to a certain extent in practical fabrication from the viewpoint of PBG formation; and the filling ratio or equivalently the light threshold selection may also be done more easily. We believe that these results are helpful in both theory and practice of HL.

Acknowledgment


This work is supported by the National Natural Science Foundation (64077005) and the Doctoral Program Foundation of Ministry of Education (20020422047), China.

Holographic assembly of quasicrystalline photonic heterostructures

Yael Roichman
David G. Grier
Department of Physics and Center for Soft Matter Research, New York University, New York, NY 10003
Abstract. 
Quasicrystals have a higher degree of rotational and point-reflection symmetry than conventional crystals. As a result, quasicrystalline heterostructures fabricated from dielectric materials with micrometer-scale features exhibit interesting and useful optical properties including large photonic bandgaps in two-dimensional systems. We demonstrate the holographic assembly of two-dimensional and three-dimensional dielectric quasicrystalline heterostructures, including structures with specifically engineered defects. The highly uniform quasiperiodic arrays of optical traps used in this process also provide model aperiodic potential energy landscapes for fundamental studies of transport and phase transitions in soft condensed matter systems.
pacs: (140.7010) Trapping; (090.1760) Computer holography; (120.4610) Optical fabrication

Figure 1. Two-dimensional colloidal quasicrystals organized with holographic optical traps. (a) 5-fold. (b) 7-fold. (c) 8-fold. (d) An octagonal quasicrystal with an embedded structured defect. The scale bar in (a) indicates 5 \mathrm{\upmu}\mathrm{m}.
Quasicrystals have long-ranged orientational order even though they lack the translational periodicity of crystals. Not limited by conventional spatial point groups, they can adopt rotational symmetries that are forbidden to crystals. The resulting large number of effective reciprocal lattice vectors endows quasicrystals' effective Brillouin zones with an unusually high degree of rotational and point inversion symmetry (1). These symmetries, in turn, facilitate the the development of photonic band gaps (PBG) (2) for light propagating through quasicrystalline dielectric heterostructures (3)(4)(5), even when the dielectric contrast among the constituent materials is low. Photonic band gaps have been realized in one- (6) and two-dimensional (7) lithographically defined quasiperiodic structures. Here we demonstrate rapid assembly of arbitrary materials into two-dimensional and three-dimensional quasicrystalline heterostructures with features suitable for obtaining interesting and potentially useful diffractive effects in infrared and visible light.

Figure 2. Four views of a rolling colloidal icosahedron. (a) 5-fold axis. (b) 2-fold axis. (c) 5-fold axis. (d) Midplane. Scale bar indicates 5 \mathrm{\upmu}\mathrm{m}. The complete assembly and rolling process is shown in the associated movie.
Our approach is based on the holographic optical trapping technique (8)(9)(10) in which computer-generated holograms are projected through a high-numerical-aperture microscope objective lens to create large three-dimensional arrays of optical traps. In our implementation, light at 532 \mathrm{n}\mathrm{m}from a frequency-doubled diode-pumped solid state laser (Coherent Verdi) is imprinted with phase-only holograms using a liquid crystal spatial light modulator (SLM) (Hamamatsu X8267 PPM). The modified laser beam is relayed to the input pupil of a 100\times NA 1.4 SPlan Apo oil immersion objective mounted in an inverted optical microscope (Nikon TE-2000U), which focuses it into traps. The same objective lens is used to form images of trapped objects, using the microscope's conventional imaging train (10).
We used this system to organize colloidal silica microspheres 1.53 \mathrm{\upmu}\mathrm{m} in diameter (Duke Scientific Lot 5238) dispersed in an aqueous solution of 180:12:1 (wt/wt) acrylamide, N,N^{\prime}-methylenebisacrylamide and diethoxyacetophenone (all Aldrich electrophoresis grade). This solution rapidly photopolymerizes into a transparent polyacrylamide hydrogel under ultraviolet illumination, and is stable otherwise. Fluid dispersions were imbibed into 30 \mathrm{\upmu}\mathrm{m} thick slit pores formed by bonding the edges of #1 coverslips to the faces of glass microscope slides. The sealed samples were then mounted on the microscope's stage for processing and analysis.

Figure 3. Holographic assembly of a three-dimensional colloidal quasicrystal. (a) Colloidal particles trapped in a two-dimensional projection of a three-dimensional icosahedral quasicrystalline lattice. (b) Particles displaced into the fully three-dimensional configuration. The shaded region identifies one embedded icosahedron. (c) Reducing the lattice constant creates a compact three-dimensional quasicrystal. (d) Optical diffraction pattern showing ten-fold symmetric peaks. The three-dimensional assembly process is shown in the associated movie.
Silica spheres are roughly twice as dense as water and sediment rapidly into a monolayer above the coverslip. A dilute layer of spheres is readily organized by holographic optical tweezers into arbitrary two-dimensional configurations, including the quasicrystalline examples in Fig. 1. Figures 1(a), (b) and (c) show planar pentagonal, heptagonal and octagonal quasicrystalline domains (11), respectively, each consisting of more than 100 particles. Highlighted spheres emphasize each domain's symmetry. These structures all have been shown to act as two-dimensional PBG materials in microfabricated arrays of posts and holes (3)(12)(13)(14). As a soft fabrication technique, holographic assembly requires substantially less processing than conventional methods such as electron-beam lithography, and can be applied to a wider range of materials. Unlike complementary optical fabrication techniques such as multiple-beam holographic photopolymerization (14)(15)(16)(17), assembly with holographic optical traps lends itself to creating nonuniform architectures with specifically engineered features, such as the channel embedded in the octagonal domain in Fig. 1(d). Similar structures of comparable dimensions created lithographically have been shown to act as narrow-band waveguides and frequency-selective filters for visible light (12)(13)(18)(19).
Holographic trapping's ability to assemble free-form heterostructures extends also to three dimensions. The sequence of images of a rolling icosahedron in Fig. 2 shows how the colloidal spheres' appearance changes with distance from the focal plane. This sequence also recalls earlier reports (20)(21) that holographic traps can successfully organize spheres into vertical stacks along the optical axis, while maintaining one sphere in each trap.
The icosahedron itself is the fundamental building block of a class of three-dimensional quasicrystals, such as the example in Fig. 3. Building upon our earlier work on holographic assembly (22), we construct a three-dimensional colloidal quasicrystal by first gathering the appropriate number of sedimented spheres into a two-dimensional arrangement corresponding to the planar projection of the planned structure, Fig. 3(a). Next, we translate the spheres along the optical axis to their final three-dimensional coordinates in the quasicrystalline domain, as shown in Fig. 3(b). One icosahedral unit is highlighted in Figs. 3(a) and (b) to clarify this process. Finally, the separation between the traps is decreased in Fig. 3(c) to create an optically dense structure. This particular domain consists of 173 spheres in roughly 7 layers, with typical inter-particle separations of 3 \mathrm{\upmu}\mathrm{m}.
The completed structure was gelled and its optical diffraction pattern recorded at a wavelength of 632 \mathrm{n}\mathrm{m}by illuminating the sample with a collimated beam from a HeNe laser, collecting the diffracted light with the microscope's objective lens and projecting it onto a charge-coupled device (CCD) camera with a Bertrand lens. The well-defined diffraction spots clearly reflect the quasicrystal's five-fold rotational symmetry in the projected plane.
Quasicrystalline heterostructures are likely to be particularly well suited for photonic bandgap applications because their effective Brillouin zones are more nearly spherical than those for photonic crystals and thus optimize the overlap of local bandgaps (23). This property has been exploited in lithographically defined two-dimensional devices with micrometer-scale features. Recently, the photonic band structure of a centimeter-scale three-dimensional icosahedral quasicrystal was measured in the microwave region and shown to feature prominent gaps at the effective Brillouin zone edge (23).
Although photonic band structure calculations generally are lacking for dielectric quasicrystals, the structure in Fig. 3 is similar to the polymeric quasicrystalline lattice in Ref. (23) and has a comparable mismatch in dielectric constant. It therefore should have comparable optical properties at appropriately rescaled wavelengths in the infrared and visible bands. Moreover, we expect bandgaps to be more pronounced in the holographically defined colloidal quasicrystal because of its larger volume fraction. These benefits all could be further enhanced by drying the gel at the triple point of water to further increase both the volume fraction and the dielectric constant mismatch.
Sequential assembly starting from a planar starting configuration ensures that there is one particle at each vertex of the quasicrystal. A similar outcome could be achieved in a system with microfluidic sample control without this initial step, and indeed without imaging. Larger structures then can be fabricated by assembling smaller domains, fixing them through spatially-resolved polymerization (24) or by initiating inter-particle bonds, and then combining the sub-assemblies, in much the same spirit as shoot-and-step lithography commonly used for two-dimensional semiconductor devices.
The use of monodisperse silica spheres for this demonstration was dictated by their commercial availability, low optical absorption, and high density. Holographic assembly lends itself just as easily to other materials, and so can be tailored to particular applications. Deterministic organization of disparate components under holographic control could be used to embed gain media in photonic bandgap cavities, to install materials with nonlinear optical properties within waveguides to form switches, and to create domains with distinct chemical functionalization. Distinctly engineered domains can be combined into larger heterostructures through sequential assembly, with larger-scale structures lending themselves to conventional stamping technologies. In all cases, this adaptable soft fabrication process can be directed toward creating mechanically and environmentally stable materials that can be integrated readily into larger systems.
Beyond the immediate application of holographic trapping to fabricating quasicrystalline materials, the ability to create and continuously optimize such structures provides new opportunities for studying the dynamics (10) and statistical mechanics (25) of colloidal quasicrystals. The optically generated quasiperiodic potential energy landscapes developed for this study also should provide a flexible model system for experimental studies of transport (26) through aperiodically modulated environments.
We are grateful to Paul Steinhardt, Paul Chaikin and Weining Man for illuminating conversations. Support was provided by the National Science Foundation through Grant Number DMR-0451589.


Multiple-beam holography has been widely used for the realization of photonic quasicrystals with
high rotational symmetries not achievable by the conventional periodic crystals. Accurate control of
the properties of the interfering beams is necessary to provide photonic band-gap structures. Here we
show, by FDTD simulations of the transmission spectra of 8-fold quasiperiodic structures, how the
geometric tiling of the structure affects the presence and properties of the photonic band-gap for low
refractive index contrasts. Hence, we show an interesting approach to the fabrication of photonic
quasicrystals based on the use of a programmable Spatial Light Modulator encoding ComputerGenerated
Holograms, that permits an accurate control of the writing pattern with almost no
limitations in the pattern design. Using this single-beam technique we fabricated quasiperiodic
structures with high rotational symmetries and different geometries of the tiling, demonstrating the
great versatility of our technique.
Keywords: computer holography, photonic band-gap materials, microstructure fabrication
________________________________________________________________________________
1. Introduction
Quasicrystals are structures exhibiting long-range
aperiodic order and rotational symmetry [1-3].
Mesoscale quasicrystals may possess photonic
bandgaps (PBGs) [4-6] that are more isotropic than
in conventional photonic and, hence, the PBG
becomes more spherical leading to interesting
properties of light transmission [7], wave guiding
and localization [8], increasing the flexibility of
these materials for many photonic applications, also
due to the possible presence of many non equivalent
defect sites. To construct two-dimensional (2D) or
three-dimensional (3D) quasicrystals is a very
difficult task. Previously used to realize periodic
photonic crystals [9,10], holographic lithography
was recently proposed and used to realize
quasicrystals too at the mesoscale [11-15]. The
holographic lithography is based on the interference
pattern of many coherent light beams, that are
usually obtained by splitting a single laser beam by
suitable grating [11,13], prism [15] or dielectric
beam splitters [16], in a single or multiple exposure
process [12]. Realizing a quasicrystal structure,
exhibiting N-fold symmetry [14], requires to control
the amplitude and phase of N interfering laser
beams, leading to several difficulties. By controlling
the relative phases of the interfering beams,
different geometries in the tiling of the dielectric
medium (typically rods, in a binary pattern,
corresponding to the maxima positions of the light
distribution) are achievable [15]. Moreover,
aperiodic structures cannot be realized even in
principle by multiple-beam interference, as for
instance the Thue-Morse structure, defined by
recursive substitutional sequences [17-20]. Onedimensional
and two-dimensional Thue-Morse
structures are known to exhibit PGB with interesting
omnidirectional reflectance [21-24].
In this communication, we demonstrate the
importance of the tiling geometry in the
arrangement of a quasicrystal structure. We show,
by FDTD (Finite Difference Time Domain)
simulations of the transmission spectra in several 8-
fold quasiperiodic patterns, the influence of different 
2
building tile geometries on the photonic band-gap
for low refractive index contrasts. This demonstrates
the importance of an accurate control of the writing
pattern to produce feasible photonic band-gap
structures with low refractive index materials. The
typical interference patterns obtained changing the
relative phases of eight interfering light beams in a
multiple-beam process (examples are shown in [15])
produce structures with important differences in the
dielectric distribution with respect to the standard
Ammann-Beenker octagonal tiling of space with
“squares” and “rhombuses” of equal side lengths a
[25,26]. We chose the octagonal case because its
building tile is easier to analyze with respect to
other structures like Penrose or dodecagonal. We
analyze and compare both the structures (octagonal
and 8-fold interferential pattern with different unittiles)
to provide a comprehension of the behavior of
the photonic band-gap with respect to the building
tile, the variation of the refractive index difference
Ī”n and the filling factor (high dielectric constant
area to overall area ratio).
Moreover, in this communication, we report the
fabrication of several 2D quasicrystals in the
mesoscale range, using a single-beam technique
based on the spatial modulation of the optical beam
by means of Spatial Light Modulator (SLM) and
Computer-Generated Hologram (CGH). This
method was discussed in details in a previous work
[27]. Here we show interesting results from the
FDTD simulations about quasiperiodic structures
that are difficult to achieve by multiple-beam
holography, but achievable with the SLM-CGH
technique. That being so, in this work we review the
most important aspects of our holographic technique
and the results already obtained. In fact, this method
permits to control with high accuracy the properties
of the required photonic structure achieving the
desired dielectric distribution. We were able to
produce, with single-beam optical setup, 2D
quasiperiodic patterns of rotational symmetry as
high as 23-fold, and aperiodic patterns that cannot
be realized using N-beam interference whatever
large may be N, as for instance the 2D Thue-Morse
pattern. These results demonstrate well the potential
of our single-beam technique. Our structures were
realized by induced photo-polymerization of liquid
crystal-polymer composites. Holographic Polymer
Dispersed Liquid Crystals (H-PDLCs) are materials
that provide good mechanical and optical properties
and can be switched by applying moderate external
electric fields [16]. Nevertheless, the SLM-CGH
technique might be applied to other kinds of
photosensitive materials.
2. FDTD analysis, results and discussion
The octagonal structure analyzed in this paper was
supposed made of dielectric rods with the AmmannBeenker
tiling of space, in air. The positions of the
cylinders of radius r are coincident with the vertices
of “squares” and “45° rhombuses” with sides of
equal length a. This structure presents a complete
PBG with a very low threshold value for the
refractive index difference (Ī”n=0.26) between high
and low dielectric materials, whereas the gap width
to midgap ratio becomes close to 5% for Ī”n=0.45
[28]. Therefore, optoelectronics devices based on
the octagonal photonic quasi-crystal (PhQC)
promise to be realized in silica, a very common
telecommunication optical material, or even in soft
materials like polymer. Permitting to record largearea
photonic quasicrystals in photosensitive
materials and exploiting, typically, such substrates,
holographic lithography, therefore, represents an
important fabrication technique. The writing pattern
of light is usually obtained as multiple-beam
interference. In fact, the interference irradiance
profiles I(r), according to [11,15]
1 1
( ) exp[ ( ) ( )],
N N
mn m n m n
m n
I AA i i Ļ• Ļ• ∗
= =
r k kr = − ⋅+ − ∑ ∑ (1)
where Am, km, Ļ†m, are the amplitudes, the wave
vectors and the initial phases of the interfering
beams, respectively, give quasiperiodic distributions
of the dielectric material in the recording medium.
Depending on the threshold level of the
photosensitive matter and the exposure time, the
filling factor may have different values. Typically,
the maxima positions of the light pattern correspond
to the high dielectric regions, that, usually, can be
approximated with a structure of dielectric rods in
air (or other materials). The number N of the
interfering beams determines the order of the
rotational symmetry of the quasicrystal pattern [29].
The wave vectors of the interfering beams are given
by [15]
22 2 m (sin( )sin ,cos( )sin ,cos ), nm m
N N
Ļ€Ļ€ Ļ€ Īø Īø Īø
Ī»
k = (2)
where the km, m=(1, …, N), are oriented at angle Īø
with respect to the longitudinal z-direction, and are
equally distributed along the transverse (x, y)-plane;
n is the average refractive index of the
photosensitive mixture, and Ī» is the common
wavelength of the beams. By adjusting the
parameters in equation (1), different interference
patterns can be obtained (figure 1). By changing the
relative initial phases of the beams, different spatial
3
Figure 1. (a) Calculated quasiperiodic irradiance profile (IP) with 8-fold rotational symmetry obtained from phase
values Ļ†i = 0, for i ={1,…,8}, say, 8-fold(A) pattern; (b) calculated 8-fold IP from phase values Ļ†1 = Ļ†5 =0, Ļ†2 = Ļ†4 = Ļ†6
= Ļ†8 = Ļ€/2, Ļ†3 = Ļ†7 = Ļ€, say, 8-fold(B) pattern. (c)-(e) Calculated IPs from 9-, 10-, 12-beam interference, respectively.
arrangements of the rods in air may be realized,
depending on the resulting interferential profiles.
We compared the octagonal pattern with the 8-fold
interferential structures obtained for two different
sets of the initial phases. For the first pattern, say, 8-
fold(A), the phases were supposed to be all equal,
that is Ļ†1 = …= Ļ†8 =0, whereas for the second
pattern, say, 8-fold(B), the phases were periodically
shifted of Ļ€/2, that is Ļ†1 = Ļ†5 =0, Ļ†2 = Ļ†4 = Ļ†6 = Ļ†8 =
Ļ€/2, Ļ†3 = Ļ†7 = Ļ€.
Figure 2. (a) dielectric structure of rods from 8-fold(A)
pattern; (b) dielectric structure of rods from 8-fold(B)
pattern; (c) octagonal structure of rods with “squarerhombus”
tile of equal side lengths a.
The intensity profiles calculated for both the 8-
fold(A) and 8-fold(B) structures are shown in figure
1-(a) and 1-(b), respectively. The corresponding
patterns of rods are shown in figure 2-(a) and 2-(b),
respectively. These structures are obtained by
positioning circular dielectric rods of radius r in the
maxima of the interferential intensity pattern IP of
figure 1-(a) and 1-(b), respectively. In figure 2-(c),
the octagonal structure of rods with the AmmannBeenker
tiling of space is depicted. The side length
of the unit-tile of space is a for the octagonal pattern
in 2-(c). Due to the non-geometric procedure of
building the structures depicted in figures 2-(a) and
2-(b), we found more convenient to define as
characteristic length of the patterns a new
parameter, that is the average distance ad between
neighbouring rods along the x-direction.
The 2D finite difference time domain (FDTD)
method with uniaxial perfectly matched layer
(PML) boundary conditions was used in all
simulations. The photonic quasi-crystals examined
were non-periodic in the translational direction and
the supercell approximation, necessary in these
cases, required very long computational time [30].
The FDTD technique, instead, was faster and very
accurate. We employed this approach to obtain
transmission information, through the x-y plane
(figure 2), as a function of propagation direction,
wavelength and polarization. A Gaussian time-pulse
excitation was placed in several points of the
structure (in different simulations). The pulse was
wide enough in frequency domain to cover the range
of frequencies of interest. Several detectors were
placed in particular positions and allowed us to store
field components. Their positions were chosen to
cover the angular range related to the 8-fold
rotational symmetry (45°) and the mirror symmetry
with respect to a line of 22.5° in each 45° sector
[31], with an angular separation from 5° to 15°.
After a sufficiently large time of calculation, the
field was Fourier-transformed to calculate the
transmission spectrum with very high frequency
resolution. The corresponding wavelength range
was (0.1-6.0)Ī¼m with a resolution of Ī“=5.0×10-4
Ī¼m.
The transmission coefficient through the system was
calculated for different values of the dielectric
constant of the rods in air, for both polarization TM
(electric field Ez parallel to the rod axis) and TE
(magnetic field Hz parallel to the rod axis) [5], for
each detector, that is for different propagation
directions of the time-pulse excitation. The
discretization grid provided a minimum of 100 grid
points per free space wavelength. The transmission
coefficient, normalized with respect to the incident
power of the source, was calculated as a function of
the filling fraction r/ad, given as the rod radius to
average distance ratio. The parameter r/ad is related
to the filling factor and it was varied in order to
maximize the gap width of the octagonal structure
for fixed refractive index difference Ī”n=0.65. We
found that interesting values were in the range r/ad ≈
(0.15 – 0.35), with a maximum band width at r/ad ≈
0.24 (r/a ≈ 0.34). The value of r/ad was held fixed in
all simulations analyzed here for comparison.
Figure 3. Transmission spectra (TM polarization) of the
octagonal pattern (Ammann-Beenker tiling) for several
values of the refractive index difference Ī”n.
The data collected from the detectors placed at
different positions and angular orientations had the
same overall shape for all the transmission spectra
we investigated here, demonstrating the isotropy of
the structures with respect to the propagation
direction of the excitation source and, hence, the
existence of a complete photonic band-gap.
In figure 3, the transmission spectra of the
octagonal pattern of figure 2-(c) is shown for TM
polarization as a function of the refractive index
difference. The spectra are shifted in the vertical
direction to permit comparison. We see that an
index difference of Ī”n ≈ 0.3 is needed to form the
band-gap at the free space mid-gap wavelength
Ī»m=1.62Ī¼m (corresponding to the tile length
a=0.93Ī¼m and the average distance ad=1.0Ī¼m), with
a gap width to midgap ratio Ī”Ī»/Ī»m of 0.3%. The
PBG had Ī”Ī»/Ī»m=6.3% at Ī»m=1.74Ī¼m for Ī”n=0.5
and increased up to Ī”Ī»/Ī»m=22.6% at Ī»m=2.21Ī¼m for
 Ī”n=1.0 (not shown in figure 3). The mid-gap free
space wavelength Ī»m varied with Ī”n because the
increasing of the index difference corresponds to
increasing of the average refractive index.
We did not find a clear PBG in the octagonal
pattern of figure 2-(c) for the TE polarization. Our
simulations showed, in fact, only a very narrow
PBG with a relative width <1% for Ī”n=0.9 around
1.48Ī¼m, whereas in other regions of the spectra the
possible presence of a band-gap was completely
hidden by structures of multiple peaks that were
related, probably, to localized modes (not shown
here). Moreover, in the case of the TE polarization,
the data collected from the detectors placed at
different angular orientations demonstrated a
dependence from the propagation direction.
In figure 4-(a) and 4-(b), respectively, the
transmission coefficients in TM polarization of the
8-fold(A) and 8-fold(B) patterns (see figure 2a and
2b) are shown, for a refractive index difference
Ī”n=1.5. The two structures presented important
differences. The pattern (A) had a PBG of
Ī”Ī»/Ī»m=19% around Ī»m=3.75Ī¼m, whereas the
pattern (B) had two band-gaps, one with a relative
width Ī”Ī»/Ī»m=13% at Ī»m=1.93Ī¼m and the other of
12% at 3.30Ī¼m. We analyzed the same structures
for several values of the index contrast and TM
polarization. We found that, in comparison with the
octagonal structure, an index contrast larger of about
a factor two was needed to open a PBG of
comparable width to mid-gap ratio. The pattern (A)
had a larger band-gap, but for longer wavelengths
(Ī»m=3.75Ī¼m) with respect to the first band-gap of
the pattern (B) and with respect to the PBG of the
octagonal pattern of figure 2-(c), that had Ī»m=2.6Ī¼m
for Ī”n=1.5 (not shown in figure 3). On the other
hand, the 8-fold(B) pattern, even with a smaller
PBG, could work at shorter wavelengths
(Ī»m=1.93Ī¼m). By changing the filling fraction the
transmission spectra preserved the overall shape.
Transmission (TM)
Only a shift in the mid-gap wavelength was
observed.
Figure 4. (a) Transmission spectrum (TM polarization)
for the 8-fold(A) pattern with refractive index difference
Ī”n=1.5; (b) Transmission spectrum (TM polarization) for
the 8-fold(B) pattern with Ī”n=1.5.
Very interesting are the results related to the
calculation of the transmission spectra for the TE
polarization. In figure 5-(a) and 5-(b), the
transmission information in TE polarization related
to the 8-fold(B) pattern, calculated supposing a
refractive index difference Ī”n=0.6 and Ī”n=0.4,
respectively, are shown. Two PBGs, independent
from the propagation direction, were present, one
for Ī”n=0.4 with a relative width Ī”Ī»/Ī»m=32% at the
mid-gap free space wavelength Ī»m=1.25Ī¼m and the
other for Ī”n=0.6 with a relative width
Ī”Ī»/Ī»m=31% at Ī»m=1.30Ī¼m: very short wavelengths
in both cases compared to the other patterns. The
peaks around 1.3Ī¼m were probably due the
existence of localized modes. With the increasing of
the refractive index difference only a shift in the
mid-gap wavelength was observed, up to 2.25Ī¼m
for Ī”n=1.6. The 8-fold(A) pattern, on the other
hand, did not present low index PBG for TE
polarization (not shown here) as is also for the
octagonal structure with the Ammann-Beenker
tiling.
The behaviour of the 8-fold(B) pattern
represents, in our opinion, a surprising result
promising the implementation of reliable low index
contrast PhQC devices in polymeric substrates.
Figure 5. (a) Transmission spectrum (TE polarization)
for the 8-fold(B) pattern with refractive index difference
Ī”n=0.6; (b) Transmission spectrum (TE polarization) for
the 8-fold(B) pattern with Ī”n=0.4.
We analyzed also the complementary dielectric
structures of the 8-fold(A) and (B) patterns, that is
circular rods of the low index material (air)
embedded in the high index substrate. Also in this
case no band-gap was observable for the low
refractive index contrasts examined. Nevertheless,
the simulations were performed only for TM
polarization in this case.
3. Holographic patterning and fabrication
As demonstrated in the previous section, accurate
control of the dielectric arrangement is fundamental
to provide the desired low index PBG in
quasicrystal structures. The interferential structure
designated as 8-fold(B) might be a good candidate
for this purpose. For this reason, in this section, we
review the Computer-Generated Holography
technique, a single-beam technique developed in a
previous work [27] that permits to fabricate almost
any kind of structure, interference-based or not, with
high reproducibility and very accurate control of the
light intensity distribution in the writing process.
Computer-generated holography is an attractive
technique that allows to create, with almost any
design, two-dimensional, even three-dimensional,
Transmission (TM)
spatial distributions of the optical beam intensity by
controlling the phase profile of a laser impinging on
a Diffractive Optical Element (DOE) [32,33]. In our
work, we implemented the DOE by means of a
programmable liquid crystal SLM that can encode
the CGH in its LC-display. While in the standard
multi-beam holography is quite difficult to work
with a large number of beams [29], with the CGHSLM
it is possible to encode the desired spatial
distribution in a single beam. Although limited from
the pixel resolution of the SLM, our CGH-SLM
technique alleviates the great difficulty in
maintaining phase coherence of the interfering
beams between subsequent exposures, providing
high reproducibility of the resulting structures.
Moreover, the real-time switching from one pattern
to the other is easily accomplished without changing
the optical alignment. We adopted a liquid crystal
spatial light modulator HoloEye LC-R 3000,
permitting intensity images of 256 grey levels with a
maximum resolution of 1200×1920 pixels. The
CGHs corresponding to the optical phase profiles to
be added to the incident beam have been addressed
to the SLM via computer. By using an iterative
algorithm it is possible to create the desired
irradiance pattern in the Fourier plane of the CGH
[32,34,35]. Alternatively, we encoded the desired
intensity pattern directly into a phase-only CGH
[36], with a “direct imaging” method, discussed in
details in [27].
The spatial light modulator LC-R 3000 (for
visible light at Ī»=532nm) had a pixel pitch of
9.5Āµm. We used a mixture of the monomer
dipentaerythrol-hydroxyl-penta-acrylate DPHPA
(60.0% w/w), the liquid crystal BLO38 by Merck
(30.0% w/w), the cross-linking stabilizer monomer
N-vinylpyrrolidinone (9.2% w/w) and a mixture of
the photoinitiator Rose Bengal (0.3% w/w) and the
co-initiator N-phenylglycine (0.5% w/w) [34,37,38].
The polymer had a refractive index np=1.530,
whereas the LC BLO38 had an ordinary refractive
index no=1.527 and an extraordinary refractive
index ne=1.799. The average refractive index was
estimated to be n∼1.57. Although the refractive
index contrast achievable with such a mixture,
usually, is Ī”n ∼ 0.2 - 0.3, the liquid crystal content
can be removed to increase the index difference up
to ∼ 0.5. Different mixtures or materials can be used
in order to increase the amplitude of the dielectric
modulation.
Here we review the results obtained with our
technique [27]. We realized quasi-periodic patterns
with 8-fold rotational symmetry, in particular the 8-
fold(A) and 8-fold(B) pattern described in the
previous section (see figure 1a-2a and 1b-2b,
respectively). Furthermore, we were able to realize
other two-dimensional quasicrystal structures with
9-, 10-, 12-, 17-, 23-fold rotational symmetry, and
the aperiodic structure based on the Thue-Morse
sequence [17]. Figures from 6-(a) to 6-(h) refer to
these structures [39]. Three insets are shown in each
figure, consisting of: the calculated Irradiance
Profile (IP) sent to our SLM and then imaged at the
sample position; the calculated 2D Fourier
Transform (FT) of this irradiance pattern; and the
experimental Diffraction Pattern (DP) produced by
the written structure. These structures are twodimensional
phase gratings in which the modulation
profile of the average refractive index is not easily
accessible with the scanning electron microscopy.
The similarity between the positions and cone angle
of the spots in the calculated Fourier transform FT
and the observed DP doubtless substantiates the
presence of the index profile and the good quality of
the samples. The DPs show the expected N-fold
symmetry having N points for even symmetry and
2N points for odd symmetry. We estimated the size
d of the “self-similarity cell” (that is the selfrepeating
basic structure observable in the
calculated intensity profiles) from the magnitudes of
the basic reciprocal vectors in the diffraction
patterns, that are related to sensible lengths of the
crystal structures, such as the tile side [16,40].
Although the size d reported in figure 6 (IP insets)
was in the range between 4.8 and 8.6Āµm, increasing
with the order of the symmetry, the typical
separation between equal dielectric regions in the
fine structure of the self-similarity cell was in the
range of ∼1 - 2Āµm. In fact, we were able to realize
1D Bragg grating and 2D periodic square lattice
with a pitch of ∼1Āµm (not shown here). The scale
length of the realized structures depended on the
lateral magnification of the relay lenses [27] (which
also affected the extension of the writing area). The
achievable resolution, lastly, depends on the SLM
pixel size and on the wavelength of the writing light.
Using UV light and state-of-the-art SLM pixel size
[41], the limit in the spatial resolution could be
improved by a factor of 3-4. Anyway, the scalability
of the optical properties permits to use our
quasiperiodic structures to study their optical
behavior. Furthermore, according to the TE
transmission spectra for the 8-fold(B) pattern of
figure 5 presented in the previous section, an
average distance ad∼1.25Āµm between neighbouring
rods would be sufficient to induce a PBG at
Ī»m=1.55Ī¼m. That being so, the SLM-CGH
technique could be employed to produce photonic
quasicrystals for feasible and reliable applications. 
7
Figure 6. (a) Quasiperiodic structure 8-fold(A) with phases Ļ†i = 0, for i ={1,…,8}; (b) quasiperiodic structure 8-fold(B)
with phases Ļ†1 = Ļ†5 =0, Ļ†2 = Ļ†4 = Ļ†6 = Ļ†8 = Ļ€/2, Ļ†3 = Ļ†7 = Ļ€. (c)-(g) Quasiperiodic structure with 9-, 10-, 12-, 17-, 23-
fold rotational symmetry, respectively. (h) Two-dimensional Thue-Morse quasicrystal structure. (a)-(h) Top left inset:
calculated irradiance profile (IP); top right inset: 2D Fourier transform (FT) of the irradiance profile; bottom inset:
observed diffraction pattern (DP); d estimates the self-similarity cell size of the structures we fabricated.
4. Conclusion and discussion
In conclusion, we showed, by FDTD simulations of
the transmission spectra of 8-fold quasicrystals, the
influence of different building tile geometries on the
photonic band-gap, demonstrating the importance of
an accurate control of the writing pattern to produce
feasible photonic band-gap structures with low
refractive index materials. Besides the octagonal
pattern with the Ammann-Beenker geometric tiling,
the 8-fold(B) pattern of figure 2-(b) was found very
interesting among the interferential structures
because it permits to obtain a large PBG for an
index contrast as low as 0.4, as shown in figure 5.
For this purpose we presented a fabrication
technique that alleviates many drawbacks of the
multi-beam holographic lithography that is typically
employed for the PhQCs realization. We used
Computer-Generated Holograms to drive a liquid
crystal Spatial Light Modulator. This permits to
write arbitrary structures with simple and reliable
optical setup. The writing intensity pattern can be
modified by real time change of the CGH addressed
to the programmable SLM. Accurate control of the
pattern designs and of the tiling geometry can be
achieved without mechanical motion and optical
realignment. Together with the 8-fold quasicrystal
patterns discussed in the section 2, we were able to
obtain, with a single-beam technique, quasiperiodic
structures with unprecedented rotational symmetries
up to 17- and 23-fold and two-dimensional ThueMorse
structures too. The possibility of a full
characterization of quite interesting aperiodic
structures not achievable by multiple beam
holography even in principle (like two-dimensional
Thue-Morse), and that have been studied until now
only theoretically [24], is open. Our structures were
written into polymeric liquid crystal films, so to
permit switching by external fields [16]. However,
our holographic technique could be applied, in
principle, to any photosensitive material (e.g. with a
larger index contrast), or to produce patterned masks
and templates for use in lithography of hard
materials [9-16,29]. We expect that the SLM based
CGH technique may have a large impact on the
production of complex photonic structures.


References


[1] Levine D and Steinhartdt P J 1984 Quasicrystals: a
new class of ordered structures Phys. Rev. Lett.
53 2477-80

[2] Steinhartdt P J and Ostlund S 1987 The Physics of
Quasicrystals (River Edge, NJ: Word Scientific)

[3] Stadnik Z M 1999 Physical Properties of
Quasicrystals (New York: Springer)

[4] Zoorob M E, Charlton M D B, Parker G J,
Baumberg J J and Netti M C 2000 Complete
photonic bandgaps in 12-fold symmetric
quasicrystals Nature 404 740-3

[5] Chan Y S, Chan C T and Liu Z Y 1998 Photonic
band gaps in two-dimensional photonic
quasicrystals Phys. Rev. Lett. 80 956-9

[6] Zhang X, Zhang Z Q and Chan C T 2001 Absolute
photonic band gaps in 12-fold symmetric
photonic quasicrystals Phys. Rev. B 63
081105/1-4

[7] Cheng S S M, Li L, Chan C T and Zhang Z Q
1999 Defect and transmission properties of twodimensional
quasiperiodic photonic band-gap
systems Phys. Rev. B 59 4091-99

[8] Jin C, Cheng B, Man B, Li Z, Zhang D, Ban S and
Sun B 1999 Band gap and wave guiding effect
in a quasiperiodic photonic crystal Appl. Phys.
Lett. 75 1848-50

[9] Campbell M, Sharp D N, Harrison M T, Denning
R G and Turberfield A J 2000 Fabrication of
photonic crystals for the visible spectrum by
holographic lithography Nature 404 53-6

[10] Berger V, Gauthier-Lafaye O, Costard E 1997
Fabrication of a 2D photonic band gap by a
holographic method Electr. Lett. 33 425-6

[11] Wang X, Ng C Y, Tam W Y, Chan C T and Sheng
P 2003 Large-area two-dimensional mesoscale
quasicrystals Adv. Mater. 15 1526-28

[12] Gauthier R C Ivanov A 2004 Production of
quasicrystal template patterns using a dual beam
multiple exposure technique Opt. Express 12
990-1003

[13] Wang X, Xu J, Lee J C W, Pang Y K, Tam W Y,
Chan C T and Sheng P 2006 Realization of
optical periodic quasicrystals using holographic
lithography Appl. Phys. Lett. 88 051901/1-3

[14] Gorkhali S P, Qi J and Crawford G P 2005
Electrically switchable mesoscale Penrose
quasicrystal structure Appl. Phys. Lett. 86
011110/1-3

[15] Yang Y, Zhang S and Wang G P 2006 Fabrication
of two-dimensional metallodielectric
quasicrystals by single-beam holography Appl.
Phyis. Lett. 88 251104/1-3

[16] Gorkhali S P, Qi J and Crawford G P 2006
Switchable quasicrystal structures with five-,
seven-, and ninefold symmetries J. Opt. Soc.
Am. B 23 149-158

[17] MaciĆ  E 2006 The role of aperiodic order in
science and technology Rep. Prog. Phys. 69 397-
441

[18] Axel F and Terauchi H 1991 High-resolution Xray-diffraction
spectra of Thue-Morse GaAsAlAs
heterostructures: towards a novel
description of disorder Phys. Rev. Lett. 66 2223-
26

[19] Axel F and Terauchi H 1994 Axel and Terauchi
reply Phys. Rev. Lett. 73 1308

[20] Kolar M 1994 High-resolution X-ray-diffraction
spectra of Thue-Morse GaAs-AlAs
heterostructures Phys. Rev. Lett. 73 1307

[21] Dal Negro L, Stolfi M, Yi Y, Michel J, Duan X,
Kimerling L C, LeBlanc J and Haavisto J 2004
Photon band gap properties and omnidirectional
reflectance in Si/SiO2 Thue-Morse quasicrystals
Appl. Phys. Lett. 84 5186-88

[22] Lee H Y and Nam G Y 2006 Realization of
ultrawide omnidirectional photonic band gap in
multiple one-dimensional photonic crystals J.
Appl. Phys. 100 083501/1-5

[23] Lei H, Chen J, Nouet G, Feng S, Gong Q and Jiang
X 2007 Photonic band gap structures in the
Thue-Morse lattice Phys. Rev B 75 205109/1-10

[24] Moretti L and Mocella V 2007 Two-dimensional
photonic aperiodic crystals based on ThueMorse
sequence Opt. Express 15 15314-23

[25] Zhou G and Gu M 2007 Photonic band gaps and
planar cavity of two-dimensional eigthfold
symmetric void-channel photonic quasicrystal
Appl. Phys. Lett. 90 201111/1-3

[26] Socolar J E S 1989 Simple octagonal and
dodecagonal quasicrystals Phys. Rev. B 39
10519-10551

[27] Zito G, Piccirillo B, Santamato E, Marino A,
Tkachenko V and Abbate G 2008 Twodimensional
photonic quasicrystals by single
beam computer-generated holography Opt.
Express 16 5164-70

[28] Romero-Vivas J, Chigrin D N, Lavrinenko A V
and Sotomayor Torres C M 2005 Resonant adddrop
filter based on a photonic quasicrystal Opt.
Express 13 826-35

[29] Mao W, Liang G, Zou H, Zhang R, Wang H and
Zeng Z 2006 Design and fabrication of twodimensional
holographic photonic quasi crystals
with high-order symmetries J. Opt. Soc. Am. B
23 2046-50

[30] Gauthier R C and Mnaymneh K 2005 Photonic
band gap properties of 12-fold quasi-crystal
determined through FDTD analysis Opt. Express
13 1985-98

[31] Hase M, Miyazaki H, Egashira M, Sninya N,
Kojima K M and Uchida S 2002 Isotropic
photonic band gap and anisotropic structures in
transmission spectra of two-dimensional fivefold
and eightfold symmetric quasiperiodic photonic
crystals Phys. Rev. B 66 214205/1-7

[32] Soifer V A 2002 Methods for Computer Design of
Diffractive Optical Elements (New York: John
Wiley & Sons, Inc.)

[33] Lee G, Song S H, Oh C H and Kim P S 2004
Arbitrary structuring of two-dimensional
photonic crystals by use of phase-only Fourier
gratings Opt. Lett. 29 2539-41

[34] Zito G, Piccirillo B, Santamato E, Marino A,
Tkachenko V and Abbate G 2007 Computergenerated
holographic gratings in soft matter
Mol. Cryst. Liq. Cryst. 465 371-8

[35] Goodman J W 1996 Introduction to Fourier Optics
(New York: McGraw-Hill) 

[36] Davis J A, Cottrell D M, Campos J, Yzuel M J and
Moreno I 1999 Encoding amplitude information
onto phase-only filters Appl. Opt. 38 5004-13
[37] Sutherland R L, Natarajan L V, Tondiglia V P and
Bunning T J 1993 Bragg gratings in an acrylate
polymer consisting of periodic polymerdispersed
liquid-crystal planes Chem. Mater. 5
1533-38

[38] Vita F, Marino A, Tkachenko V, Abbate G,
Lucchetta E D, Criante L and Simoni F 2005
Visible and near-infrared characterization and
modeling of nanosized holographic-polymerdispersed
liquid crystal gratings Phys. Rev. E 72
011702/1-8

[39] Figures 6(a)-(h) are taken from figures 1 and 2 of
our previous work [27] with the permission of
the publisher

[40] Kaliteevski M A, Brand S, Abram R A, Krauss T
F, Millar P and De La Rue R M 2001 Diffraction
and transmission of light in low-refractive index
Penrose-tiled photonic quasicrystals J. Phys.:
Condens. Matter 13 10459-70

[41] Apter B, David Y, Baal-Zedaka I and Efron U
2007 Experimental study and computer
simulation of ultra-small-pixel liquid crystal
device 11th Meeting on Optical Engineering and
Science (Israel, Tel Aviv) 
x
x




I believe this material could have been part of the technology used to place "fake airplanes" in the sky on 9/11.

- Christian Hampton - Wolf Clan Media

Comments

Popular posts from this blog

9/11 - Phantom Warfare

Energy & Consciousness I'd like to thank Mark Conlon for all his research.  Which made writing this article possible.   There is an overwhelming amount of  evidence that suggests that directed free energy technology was used on 9/11  to destroy the World Trade Center.  Neither bombs or fire can explain the clearly visible evidence in this 3-minute clip... Before you continue reading this article, if you haven't already done so, then please take the time to study this evidence  for yourself.  Phantom Jets  There is also  evidence that suggests the 'airplanes' on 9/11 were actually a product of psychological warfare  through the use of  holography  or some type of  optical illusion  and the 'airplane damage' was caused by a form of directed energy as well.  We have already established through the work of many researchers that the airplanes on 9/11 could not have been the 4 commercial aircraft that we were to

9/11 - Ghost Gun UA-175

Ghost Gun UA175 Holograms, Whistleblowers and the 9/11 Media Hoax Revised 9/11/2005 |  Updates @ bottom | w/radio show Re-posted 4/28/2018 With minor edits and adjustments made by:  Wolf Clan Media I do no agree with the conclusion of "pre-planted explosives" creating the airplane shaped holes in the Twin Towers.  I strongly feel this was done using un-conventional technology. See my article on this ---> http://wolfclanmedia-research.blogspot.com/2018/03/911-phantom-warfare.html Most 9/11 (no-plane) researchers promote falsehoods about the video & photographic evidence from that day. Some highlights of research from 2005 which makes some very great observations and helps to prove that an object was in the sky on 9/11.  Many links are now broken :( Unfortunately, the UA175 illusion (WTC2, morning of 9-11-1) was not 100% successful.   Foreword (by Dr. Stefan Grossmann): I have been asked to describe what the name ‘Ghost Gun’ (as used i

9/11 - Optical Illusions

The Word 'Holograms' & Perception Management   I've included web links for most of my statements. (numerical order) Many documents are available as well. chris@wolfclanentertainment.com Section 1     I'd like to discuss evidence of a technology that exist . This technology has been publicly known since at least 1971.   (1)  However design started obviously much earlier.  (2)  For ordinary citizens this technology is kept in the realm of Science Fiction . For some strange reason we decide to believe that's all it has, & ever will be...  I'm referring to a specific phenomenon, but the public perception of it's reality is not unique to other advanced technologies.  (3)  In other words, people feel this way about many strange things that do indeed exist.  (4)  Obviously there is a psychological aspect to this. Otherwise such words as 'Holograms' would not be so laughable when discussed publicly.  While we are briefly on the subject of