121
Research Collection Doctoral Thesis Phononic quasicrystals Author(s): Sutter-Widmer, Daniel Publication Date: 2007 Permanent Link: https://doi.org/10.3929/ethz-a-005399971 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

Phononic Quasicrystals

Embed Size (px)

Citation preview

Research Collection

Doctoral Thesis

Phononic quasicrystals

Author(s): Sutter-Widmer, Daniel

Publication Date: 2007

Permanent Link: https://doi.org/10.3929/ethz-a-005399971

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Diss. ETH Nr. 17174

Phononic Quasicrystals

A dissertation submitted to

ETH ZÜRICH

for the degree of

Doctor of Sciences

presented by

DANIEL SUTTER-WIDMER

Dipl. Ing. ETH

born 12.05.1978

citizen of Bretzwil (BL)

Accepted on the recommendation of

Prof. Dr. W. Steurer, examiner

Dr. U.G. Grimm, co-examiner

Dr. J.O. Vasseur, co-examiner

2007

Contents

1 Introduction 1

1.1 Phononic crystals 2

1.1.1 Phononic crystals in a nutshell 2

1.1.2 Gaps happen - where and why? 3

1.2 Non-periodic Structures 4

1.2.1 Aspects of quasiperiodic structures relevant for this thesis 4

1.2.2 Other aperiodic structures of interest 5

2 Phononic quasicrystals 6

2.1 Overview of structures and literature review 6

2.1.1 The peculiarities of quasiperiodic order 8

2.1.2 ID QPTC's and QPNC's 12

2.1.3 2D and 3D QPTC's and QPNC's 20

2.1.4 Applications 23

2.1.5 Fabrication of QPTC's and QPNC's 24

2.2 QPNC's and single scatterer resonance states (Article 4) 26

2.3 QPNC's anc Bragg scattering (Article 5) 38

2.4 Icosahedral phononic quasicrystals 65

2.5 QPNC's and clusters (Article 2) 72

3 Synthesis 78

A Theory of elastic waves 80

A.l The wave-equation 80

A.2 Scattering behavior of single rods 83

A.3 Transmittance and attenuation regimes 86

B Computational methods used 92

B.l The plane wave expansion method 92

B.2 The finite difference time domain approximation 93

B.3 The multiple scattering method 94

C Systematic description of quasiperiodic structures 96

Seite Leer /

Blank leaf

Abstract

In the present thesis the suitability of quasiperiodic structures for phononic crystals

(PNC's) was studied. Periodic and quasiperiodic PNC's were manufactured and char¬

acterized by ultrasonic transmission spectroscopy. The experimental work was accompa¬

nied and extended by numerical simulations using different techniques. The influence of

quasiperiodic structures was investigated separately for the two fundamental regimes of

scattering of the acoustic waves in PNC's, namely resonance scattering and Bragg scat¬

tering. For both scattering regimes a piocedure for the prediction of the existence of

isotropic band gaps is proposed.

For 2D phononic crystals which consist of scattering objects (rods) with strong reso¬

nance states in the frequency range of interest quasiperiodic structures appear to be very

favorable in all respects. The positions of the gaps follow strictly the resonance frequen¬

cies of the single rods. The isotropy of these gaps can be increased by using quasiperiodic

structures with higher rotational symmetries. Isotropy can be further enhanced by ad¬

justing the; symmetries of the fields eradiated from rods at resonance frequencies to the

symmetry of the structure by changing the shape of the scatterers (e.g., polygonal or

star-shaped rods). In higher symmetric structures the edges of the bands and gaps are

less sharply defined as compared to those of periodic PNC's. The existence of omnidirec¬

tional band gaps of such a quasiperiodic PNC (QPNC) can thus be predicted based on

knowledge; of the resonance states of the single rods.

In 2D phononic crystals in which the band gap formation evolves due to Bragg scat¬

tering, on the other hand, the positions of the band gaps are essentially determined by the

structure and indicated by strong Bragg peaks in the Fourier spectrum of the impedance

distribution of a PNC. Bragg scattering QPNC's can have spectral properties which re¬

semble more those of disordered PNC's. An approach based on the periodic average

structure (PAS) of the quasiperiodic tilings is proposed to explain this. A QPNC can be

considered as a PNC which deviates from the PAS displacively and substitutionally. The

degree of deviation from the periodic structure can be used to characterize the different

quasiperiodic tilings. The octagonal structure, for instance, deviates astonishingly little

from its PAS and the corresponding QPNC transmits similar to the PNC with the PAS.

Tilings which differ stronger from their PAS were found to show no clear bands and gaps

in their transmission spectra. The existence of omnidirectional band gaps of such a QPNC

can thus be predicted based on knowledge of the periodic average structure.

These approaches also apply to 3D QPNC's. The generally lower filling fractions as

well as the reduced tendency for wave localization lead to a clearer formation of bands

and gaps in the transmission spectra. This makes Bragg scattering 3D QPNC's more

similar to their periodic counter parts and more convenient to work with than ID or 2D

QPNC's in this regime.

m

Zusammenfassung

Die vorliegenden Arbeit befa&st sich mit der Eignung von quasipcriodischen Strukturen für

den Bau von phononischen Kristallen (PNC's). Phononische Kristalle und Quasikristalle

wurden hergestellt und mit Hilfe von Ultraschalltran&mis&ionsspektroskopie charakteri¬

siert. Die experimentelle Albeit wurde durch nummerisehe Simulationen ergänzt und er¬

weitert. Die Untersuchung des Einflusses der quasiperiodischen Strukturen ei folgte dabei

separat für PNC's mit Resonanz oder Bragg-Beugung bedington Bandlücken. Für beide

Fälle werden Ansätze vorgeschlagen, die eine apriori Beurteilung des möglichen Auftretens

von omnidirektionalen Bandlücken in quasipcriodischen phononischen Kiistallon (QPNC's)erlauben.

Bei 2D PNC's in denen die einzelnen Streukörper im untersuchten Frequenzbereich

starke Resonanzen aufweisen, wirken sich quasipeiiodi&che Strukturen weitgehend positiv

aus. Die Bandlücken treten bei den Resonanzfrequenzen der einzelnen Streukörper auf

und die hohe Rotations&ymmetrio der Anordnung führt zu einer ausgeprägten lsotiopio

der Bandlücken. Eine weitere Erhöhung dieser Isotropie kann duich ein Feinju&tieren

der Symmetrie der Streufelder erreicht weiden. Über die Form der Streukörper kann

die Symmetrie der abgestrahlten Felder an die Symmetrie der Struktur angepasst weiden.

Die Isotropie der Transmissionsspektren geht einhei mit weniger scharf definierten Kanten

der Bänder und Lücken. Die Existenz von richtungsunabhängigen Bandlücken in solchen

QPNC's kann also anhand dei Struktur und der Resonanzzustände dei einzelnen Streuer

abgeschätzt weiden.

In 2D PNC's in denen Bragg Beugung die Bandlücken erzeugt, wird die Position

der Lücken wesentlich von dei Struktur bestimmt (angedeutet durch stark1 Bragg Peaks

im Fourier Spektrum). Solche QPNC's können in Transmission änlich wie ungeordnete

Strukturen wirken und keine klaren Bänder und Lücken aufweisen. Mit Hilfe der mittleren

peiiodischen Struktur (PAS) eines quasiperiodischen Musters kann dies beurteilt werden.

Wenn man einen QPNC als PNC betrachtet, dessen Struktur der PAS der Paiketierung

entspricht und zusätzlich Fehlordnung aufweist, dann nimmt die Wahrscheinlichkeit für

das Auftreten von klaien Bandlücken im Tran&missionsspektrum ab mit zunehmendem

Mass an Fehlordnung. Die oktagonale Struktur weicht zum Beispiel von einer quadrati¬

schen PAS nur sehr geringfügig ab, und im Vergleich zu stärker aperiodischen Strukturem,

weist der oktagonale QPNC stärker ausgeprägte Bandlücken auf.

3D QPNC's sind einfacher im Umgang als ID und 2D QPNC's, speziell in Experi¬

menten. Die grundsätzlich geringeren Packungsdichten sowie die geringere Tendenz der

Wellenlokalisierung in 3D führt zu einer klareren Ausbildung von Bändern und Lücken

im Transmissionsspektrum. 3D QPNC's sind damit periodischen PNC's änlichei.

IV

Abbreviations

nD n-dimensional

n-fold «-fold Patterson symiritery

tiling qiiasiperiodic arrangement of a set of tiles or the set of vertices of it

quasilattice vertices of a qiiasiperiodic tiling

PAS Periodic average structure of a qiiasiperiodic tiling

PNC phononic crystal

PTC photonic crystal

MC meta crystal

QPNC-n r?-fold phononic quasicrystal

QPTC-n n-fold photonic quasicrystal

QMC-n ra-fold qiiasiperiodic MC

APNC phononic crystal, the structure of which corresponds to the PAS of a QPNC

transmittance transmission coefficient (as a function of frequency)

u field of displacements of elastic waves

W, tp general wave-function

[d.i]

[i]

Literature references of the bibliography at the end of the thesis

Literature references of the bibliography at the end of an article

v

Chapter 1

Introduction

Phononic crystals are a very peculiar sort of matter. They are mimicking interaction of

electron waves with real crystals on altogether different scales with altogether different

waves and altogether different materials. And yet, a similarity persists. Quasicrystals on

the other hand, arc; a very peculiar sort of matter too. They are crystals and are yet none.

The intersection of these two types of matter, phononic quasicrystals, must therefore be

(peculiar)2. And that is what they are.

We start with an outline of this thesis work, which takes oik1 peculiarity at a time.

Outline of the thesis

• Phononic crystals

— Phononic crystals in a nutshell.

- Two prominent gap forming mechanisms: Bragg gaps vs. resonance gaps.

• Quasiperiodic and related structures

- Aspects of quasiperiodicity relevant for this thesis.

— Other aperiodic structures of interest.

• Phononic Quasicrystals (articles)

The main part of the report addresses 2D and 3D phononic quasicrystals. It starts

with a comprehensive review of the; literature including a host of illustrative exam¬

ples. Then the topic is tackled based on the distinction of the fundamental types of

phononic crystals, namely resonance- and Bragg-scattering based phononic crystals.

A section on cluster-based approaches and a summing up of all relevant results of

the thesis close the chapter.

• The appendices to this work provide a collection of illustrative explanations of cer¬

tain aspects of the theory of wave propagation as well as of quasiperiodic structures.

1

1.1 Phononic crystals

1.1.1 Phononic crystals in a nutshell

Historical notes

In 1987 Yablonovitchd-163 proposed a new class of materials. A periodic distribution of two

materials of different dielectric constants, nowadays known as photonic crystals (PTC's)

or photonic bandgap materials. Less noticed but at the same time, Lakhatikia et al.d-74

published an article on elastic wave propagation in a periodic array of elastic cylinders in

a different matrix material. Also a new class of materials, nowadays known as phononic

crystals (PNC's). Both names are directly referring to the analogy of these 'meta crystals'

and reals crystals, or more precisely, electronic crystals.

Motivations

Photons or phonons encounter a periodically varying environment just as do the electrons

in the potential of a crystal lattice. Interaction of waves with a periodic potential leads

to the formation of band gaps. The optical and acoustic bands and band gaps in the

dispersion relation can be engineered just as those of electrons. The ability of thereby

controlling the propagation of waves has attracted a lot of interest.

Beyond application driven research the analogy to real crystals promised also new ac¬

cess routes to quantum physical problems in solid state physics. The formal mathematical

equivalence between the solution of the wave functions of electrons in crystals (described

by Schrödinger's equation) and the1 acoustic wave functions in PNC's (described by the

elastic wave equation) is very fascinating. The essential difference is the type of boundary

conditions involved (i.e., Dirichlet conditions suit the electron and Neuman conditions

the acoustic interface problem'122). This astonishing fact still spurs, motivates and legiti¬

mates research of a highly complex topic like; electron wave; mechanics by means of simple

acoustic experiments or calculations on easily seven orders of magnitude larger length

scales.

After the rush of investigations of perfect periodic PTC's and PNC's today certain

key areas have proved most worthy for further research, which is pursued at high pace

(see J. Dowlings database'11 for a bibliography). In photonics such a focus certainly is

on optical device physics (i.e., waveguides, add-drop-filters, etc.). In acoustics among

the most promising applications thermal barrier materials'1'21 must be mentioned. In

both fields quasiperiodic structures have become very fashionable for reasons of both

potential applications as well as their fundamentally interesting spectral properties, which

are treated in detail in chapter 2. The subject is currently covered by about 300 papers

out of tin; approximately 8700 papers on photonic crystals and the 350 papers on phononic

crystals (as of March 2007). General introductions into phononic and photonic crystals

2

can be found in a review by Sigalas et al.d 1JUand the book Photonic Crystals. Towards

Nanoscale Photonic Devices*1 M2respectively.

1.1.2 Gaps happen - where and why?

In this subsection we want to introduce two fundamental mechanisms by which most types

of band gaps can be explained. This separation is important for understanding the diffei-

ent approaches to QPNC's in the articles constituting this thesis. The two mechanisms

are both well known in solid state physics and have been used to explain electronic band

structures from the very beginning. The Bragg gap picture is best imagined as analogue

to a nearly free electron approximation and the1 resonance gap picture is best viewed as a

tight-binding system.

Bragg gaps

Bragg scattering occurs, when the wave vector, k, of the incident wave points to the

boundaries of a Brillouin zone, Bragg scattering allows reflection of waves at certain

sets of scatterers (lattice planes). The interaction of incident and reflected waves (with

wave vectors k and —k = k + G = k) enforces a splitting of the dispersion relation for

acoustic waves in a PNC, which can be explained phenomenologically in many different

ways [illustiated in Fig. 1.1(a)]. One can argue that it is merely an interference effect.

One can argue that the inter action of waves with the same wave vectors propagating in

opposite directions form standing waves ip-\ = t^m + ^scatt and -01 = V;m _ V;scatt- Standing

waves have vanishing group velocities vq — du/dk — 0 (i.e., zero net energy transfer),which implies existence of a horizontal tangent to the dispersion curve at the Bnllouin

zone boundary (i.e., a gap). One can also argue that the difference in phases of the two

waves ipz with respect to the lattice result in energy density maxima at different sites

in the unit cell/18 The two waves with equal wave vectors therefore must have different

energies (~ o>), which again describes a band gap

Bragg gaps can form only at a Brillouin zone boundary of PNC's and appears at

frequencies close to ujg ~ tt rmatnx/a0. This fiequency can, of course, be adjusted at will

by the size of the period of the structure, a0, or the wave velocity in the matrix material,

^matrix

Resonance gaps

Alternatively gaps can form if the scatteiing objects constituting a PNC have strong

resonance states in the frequency range of interest. Such PNC's are best thought of as

tight-binding systems of single resonators. The identical resonators all have the same

resonance frequencies t^s if they are independent of each other. If interaction is allowed

we can expect a iV-fold degenerate state (with N the number of scatterers) can be expected

3

(a) jt/a0 2nla0 (b)

Figure 1.1: Acoustic dispersion relation in a periodic MC (a). In periodic structures the

interaction of continuum waves (cb) with waves reflected at the Brillouin zone boundaries

(reflb) cause the formation of gaps at the zone boundaries. If there are resonance states of

the single scattering objects (b), a gap can form due to the interaction of resonance modes

(resb) with modes of the continuum band (cb). Contrary to Bragg gaps, resonance gaps

can form in periodic structures with or in aperiodic structures without zone boundaries

(b).

to form symmetrically around u;res. The with of the distribution (i.e., a band for large

N) should grow proportionally with an interaction parameter, just as in the case of

spring-coupled pendulums. This is almost exactly what happens in PNC's. Only, the

matrix medium which acts as coupling medium hosts additional wave states, which also

join the interaction process. The continuum band of the surrounding effective medium

interacts with resonance states by hybridization (mixtures of different wave states). In

the dispersion relation [see Fig. 1.1(b)) this interaction occurs at the intersection of the

flat resonance band (resb) with the linear continuum band (cb) at frequencies close to

UV,. This frequency is independent of the structure; of the PNC. To be more specific, not

even order is required, let alone periodicity - if only a minimal nearest-neighbor distance

is retainted.dUl The resonance frequencies are mainly affected by the size and the internal

wave velocity of the scattering objects.

In general, the two scattering regimes can overlap in a PNC. Resonance scattering

(Mie scattering^'88) occurring in the same frequency range as Bragg scattering favours the

formation of broad and therewith more likely omnidirectional band gaps. In this work the

band structures of PNC's an; investigated by means of transmission spectroscopy. Ranges

of band gaps are clearly indicated in such spectrum by very low transmission rates.

1.2 Non-periodic Structures

1.2.1 Aspects of quasiperiodic structures relevant for this thesis

The word quasi stems from latin and translates as so to say or equally (and should not

be mistaken for pseudo). Quasiperiodic structures can be described by the four following

aspects, which are crucial for the reminder of this work (a more constitutive description

4

follows in appendix C):

• perfect short- and long-range order

• absence of a translational period (i.e. absence of a unit cell)

• a Fourier spectrum consisting of a dense set of singular 5-peaks

• Scaling symmetry of the Fourier spectrum.

The short-range order is asserted by the fact that the structures used to construct

PNC's correspond to the vertices of a tiling of the plane (or space) with at least two

different unit tiles. A finite set of nearest-neighbor distances follows from the finite set

of tiles. A finite set of nearest-neigbor distances is important for the chemistry of real

quasicrystals because it specifies the lengths of atomic bonds but it is equally important

for the interaction of scattered near-fields in phononic quasicrystals. The long-range order

allows coherent interaction of these scattered waves. This first aspect is in focus of an

approach to understanding phononic quasicrystals in Sec. 2.2.

The second aspect introduces the major difficulty. In order to calculate or measure the

physical properties of infinite quasiperiodic systems either a large finite section or the peri¬

odic repetition of a smaller section of the quasiperiodic system can be analyzed. With in¬

creasing size of the chosen sections the properties are assumed to converge towards the ones

of the infinite system. Thus, usually there are no exact solutions and existing computer

programs can be used only for approximations (with v ry few exceptions'1'67'dfi8, d109).The third point, in return, provides some help because it indicates that certain spatial

periods arc; important for the physical properties of the quasiperiodic system. This aspect

is dealt with in an approach to understanding phononic quasicrystals in Sec. 2.3. The

definition of crystals as well as quasicrystals also evolves via the Fourier space properties.

While crystals are defined by the IUCr as structures with an essentially discrete diffraction

spectrum, quasicrystals are crystals without a spatial periodicity.

1.2.2 Other aperiodic structures of interest

Other aperiodic structures which are interesting for this work differ from quasiperiodic

structures mainly in the Fourier spectrum.

If in quasiperiodic structure a degree of periodicity can be considered (due to the fact

that the Fourier spectrum consists purely of 5-peaks) then for less periodic structures

these peaks are getting sparser and sparser and continuous parts appear. Ultimately,

in critically periodic structures the 5-peaks have transformed into a singular continuous

spectrum. For instance, a structure that has a spectrum which lacks Bragg peaks is very

interesting because of the importance of Bragg peaks for the formation of band gaps.

5

Chapter 2

Phonemic quasicrystals

2.1 Overview of structures and literature review

The aim of this overview is to work out how the peculiarities of quasiperiodic order

affect the formation of band gaps in photonic and phononic crystals but also why these

structures can be interesting with regard to potential applications of band gap materials.

The overview is combined also with an overview of the literature on both quasiperiodic

phononic an photonic: crystals. The overview focuses mainly on ID aperiodic sequences

in which the relationship between the structural nature and its effect on the propagation

of waves can be most easily demonstrated. But, also certain 2D structures are considered

but their detailed discussion can be found in subsequent sections. Promising applications

and fabrication techniques are reviewed at the end of this section.

6

Introduction

The interesting property of phononic crystals, PNC's, and photonic crystals, PTC's, is

the band gap. Such band gaps can be best studied and also best exploited in applica¬

tions if their frequency range does not depend on the crystallographic direction of wave

propagation. To give an example, a waveguide can be realized in PNC's simply by a

row of missing rods. The waveguide can only be used to direct a wave along a curved

path if the surrounding crystal is impenetrable for the wave for all possible directions

the guide happens to take. Clearly, for most periodic structures such an isotropy is not

given. It may happen that in the frequency range of a band gap in direction o there

is the center of a strong band for direction fo.dJ/11 If the gaps do not overlap in the dif¬

ferent directions, then controlling the propagation of waves is difficult because there is

always a direction in which the wave can escape from confinement. This is exactly where;

quasiperiodic structures become; interesting. Quasiperiodic order is characterized by a

discrete Fourier spectrum with arbitrarily high rotational symmetry consisting purely of

Speaks, and this despite the lack of translational symmetry. The physical properties

of systems with quasiperiodic order (e.g., band gaps) can thus become highly isotropic.

Beyond isotropy of diffraction properties, quasiperiodic structures are interesting because

of their scattering activity on multiple scales as well as the special type of wave functions,

which are interesting from a point of view of fundamental physics.

But, what really is the difference between periodic and quasiperiodic order, and how

relevant is it for practical applications in such meta-crystals (MC's)? Basically, the un¬

derlying physics (i.e., the; scattering mechanisms discussed in Sec. 1.1.2) an; exactly the

same for QMC's and MC's. This is also true for potential applications as well as the

fabrication techniques. This, justifies to focus on the structural aspects of band gap engi¬

neering. One special point must thereby be considered at all times. MC's usually consist

only of hundreds to thousands of building units. In the comparison to real intermetallic

quasicrystals in which the number of atoms is larger by almost twenty orders of magnitude

the question arises, when are MC's large enough to exhibit typical physical characteristics

of quasiperiodic systems? Or in the light of future applications even more relevant may

be the question, when are MC's large enough to inherit those characteristics of quasiperi¬

odic structures, which an1 required for optimal performance. Especially the presence of

localized waves makes transport properties of finite quasiperiodic MC's prone to severely

depend on the size of the sample.

After the following summary of general aspects of quasiperiodic patterns, the structure

of this overview follows the dimensionality of the MC's. This is almost equivalent with a

chronological ordering due to the higher complexity of experimental as well as theoretical

approaches to 2D and 3D QMC's. For all dimensions, typical representatives are discussed.

7

2.1.1 The peculiarities of quasiperiodic order

Fourier spectrum

The Fourier spectrum (kinematic diffraction pattern), P(k) reveals best the; characteristics

of ordering types (see, for instance, Baaked11 or Axel and Gratiasd9). Generally, three

terms can contribute to the Fourier spectrum

F(k) = Ppp(k) + Psc(k) + PD,(k). (2.1)

The pure point part, Ppp(k), refers to the Bragg reflections, the absolute continuous

part, Pa[ (k), is a continuous function (diffuse scattering) and the singular continuous part,

Pgc(k), is somewhere in between. It is neither continuous nor does it have Bragg peaks.

It does have peaks but these are never isolated and for increasing resolution reveal more

and more detailed subpeak structures. The integrated diffraction intensity looks like a

Cantor function (see Janner'iq). The Fourier spectrum of random(ized) structures is abso¬

lute continuous but may show broadly-peaked diffuse scattering due to local correlations.

Infinite periodic structures have a pure point Fourier spectrum. The set of diffraction vec¬

tors, k, of the Bragg peaks of a dB structure form a Fourier (Z-) module of rank n = d.

Quasiperiodic structures, such as the Fibonacci sequence or the Penrose tiling, exhibit

a pure point spectrum as well, however, the rank of their Fourier module exceeds their

dimensionality, n > d (see appendix C). A Fourier module of infinite rank characterizes

almost periodic structures such as regular fractals like the Sierpinski gasket. The peaks

in spectra with n > d densely cover the plane (fill the space) and may already appear

singular continuous. An example case for a structure with a really singular continuous

spectrum is for instance; the Time-Morse sequence and for an absolute continuous Fourier

spectrum the Rudin-Shapiro sequence can be mentioned.'141

The nomenclature; of QCs is based on their experimentally accessible diffraction sym¬

metry. The point group of tin; diffraction pattern is always centrosymmetric and symmor-

phic (i.e. apart from translations, only point group operations act as generating symmetry

operations). Consequently, all 2D A^-fold tilings with odd N, exhibit Patterson symmetry

(2N)mm, and those with TV even, the symmetry Nrnrn. The 2D Penrose tiling, for in¬

stance, has Patterson symmetry 10mm although it exhibits locally 5-fold symmetry only.

In the following, we will use the notation QPTC-/V, QPNC-/V or simply QMC-N for

quasiperioidc heterostructures with A^-fold Patterson symmetry. If in the literature the

terms 5-, 7- or 9-fold symmetry are given, we use 10-, 14- and 18-fold symmetry instead.

Real space structure, Tilings and coverings

Quasiperiodic systems are often treated as intermediate between periodicity and random¬

ness. This intermediary stems from their spectral or their macroscopic physical properties.

It must not be mistaken for a structural aspect. Quasiperiodic structures are perfectly de-

8

terministic and long-range ordered as are periodic structures. The direction dependence of

some; physical properties of quasiperiodie systems may indeed be closer to that of random

systems than to that of perioidc systems. For instance, the elasticity tensor of icosahedral

quasicrystals has only two independent coefficients like it has in amorphous materials,

while cubic crystals have three. But again, real crystal structures are never truly random

structures as a point sets can be;. Real crystal structures need not to have a minimum

distance, for instance. Although an overlap of scattering objects of a MC could be ac¬

complished (for instance, if particles partially merge by sintering), this is not generally

interesting. Such a defect would rather be treated as a substitution of a certain scattering

object by a larger one. Random point sets do not generally fulfill the Delone condition,

which periodic and quasiperiodie point sets do. The Delone condition states that the

point distribution should be uniformly discrete (minimum distance; between points) and

relatively dense (maximum hole size) (see, for instance, BaakedU). Structures may, how¬

ever, be formed by random arrangements of a set of unit tiles or deviate randomly from

an ordered structure. Randomization destroys the correlations in a structure, which is

reflected directly in the Fourier spectrum.

While for periodic tilings (lattices) a single prototile suffices to cover the plane, for

quasiperiodie tilings at least two unit tiles are required. The regular Penrose tiling, for

instance, consists of two types of tiles, a skinny and a fat rhomb. One consequence of the

larger number of tiles is the increase of different possible vertex coordinations grows with

the number of unit tiles, which can be interesting for defect creation. If the unit tiles

are arranged in a (complex) periodic way, an approximant is formed. Quasiperiodicity

can be enforced by imposing matching rules that specifiy unambiguously how a certain

tile has to be joined by surrounding tiles. Alternativly, the set of unit tiles and the

corresponding matching-rules can be transformed into a unit cluster and corresponding

overlapping rules.dM In both descriptions a relaxation of the rule of connectivity permits

randomization and disorder.

A third possibility to construct a tiling is given by the higher-dimensional approach.d5rj

A quasiperiodie structure can be generated as intersection of a r?D hypercrystal, deco¬

rated with (n — d)D atomic surfaces (occupation domains) with the dD physical space (a

pedagogically more appealing description is given in appendix C). The dimension n of the

(embedding space, V, is determined by the rank of the Fourier module of the quasiperioidc

structure. The embedding space consists of the two orthogonal subspaces, the perpendic¬

ular space, V-1-, and the physical or parallel space V". The set of n basis vectors spanning

the nD lattice not commensurate with the V. Consequently, the physical space cuts the

hypercrystal irrationally, i.e. the cut hyperplane does not contain any nD lattice point

besides the origin (see Fig. C.l). If the hypercrystal is sheared parallel to the perpendic¬

ular space, additional lattice points fall into V" and the intersection results in a periodic

structure. Such a rational approximant is not equivalent to a patch punched out of a

9

quasiperiodic tiling.

The simplest way to generate a quasiperiodic tiling is by the generalized dual-gr'id

method*-17',U7

The periodic average structure (PAS)

The periodic average structure (PAS) of a quasiperiodic tiling can be obtained by oblique

projection of the nD hypercrystal onto parallel space.dAS' d137 The nD hyperlattice is

projected onto a simple periodic lattice (see Fig. C.2). The nodes of this lattice are deco¬

rated with the projections of the atomic surfaces. Equivalcntly, if the infinite quasiperiodic

structure is projected into one unit cell of its PAS, the vertices all fall into the projected

atomic surfaces leaving the rest of the cell empty.

In principle every pair of strong Bragg peaks defines a reciprocal PAS. The significance

of these infinitely many PASs is weighted by amplitudes of the Bragg peaks defining them.

The periodicity of the PAS allows to define a Brillouin zone (BZ) as usual. The Jones zone

(JZ) or pseudo-BZ used for aperiodic crystals, is spanned by the same Bragg peaks as is

the PAS only, the pseudo BZ does not form a periodically repeatable reciprocal unit cell

but has the symmetry of the diffraction pattern. In a crude approximation, a QMC can

be seen as perturbation of its PAS. The first strong Bragg reflection, which is common to

the quasiperiodic structure and its PAS, induces tin; first band gap equally to the QPNC

and the PNC with its PAS.d145 Beyond this, the PAS allows further characterisation of

a tiling with respect to its potential in QMC's as is discussed in Sec. 2.3. The concept of

the PAS can be meaningfully applied to all structures with a Fourier module.

Properties of quasiperiodic structures relevant for wave propagation

Waves in MC's encounter a spatially modulated impedance distribution and therefore

incur multiple diffraction and refraction. The resulting interference wave fields can either

be mobile and transport energy or become localized. These phenomena are quite well

understood already for periodic and disordered periodic MC's'1,125 (see appendix A.3 for

an introduction). As discussed already in the introductory section Sec. 1.1.2 the first

gap in the transmission spectrum of a QMC is often related to the first strong Bragg

reflection.'1'61'dl5G The symmetrically equivalent MC directions along which such band

gaps appear follow the diffraction symmetry. The1 higher the; rotational symmetry the

closer to a circle is the Brillouin-zone and the more overlapping are the band gaps in

the different directions. An reasonable overlap of gaps for all directions of transmission

is therefore achieved even when the gaps are narrow. Therewith constituent materials

can be used with lower impedance contrast than for the best MC's with crystallographic

symmetry. This is particularly important in the cases of self-organized colloidal MC's,

because usually only low impedance contrast can be achieved in such systems.'1135

10

What is the equivalent to propagating Bloch waves in QMC's? Despite the fact that

mathematically the Bloch/Floquet theorem does not hold for quasiperiodic structures the

observation of the Borrmann effect [i.e., anomalous (easy) transmission of X-rays through

a perfect crystal] in icosahedral Al-Mn-Pd quasicrystalsd13, dM indicates the existence of

Bloch-like waves. For anomalous transmission, a standing wave must exist with its nodes

at the planes of highest electron densities and for quasiperiodic structures these planes are

the lattice planes of the PAS.d H0 This is also true for QMC's, and we can assume that

we have propagating waves related to the respective PAS. The broader distribution of

averaged scattering densities of the PAS compared to that of the MC may be one reason

for the slower evolution of Bragg gaps in the transmission spectra of QMC's.

If suitable parameters are found for a QPNC system the scaling symmetry as well as the

self-similarity of the; diffraction pattern are both reflected in the transmission spectra and

the band structure. This is best achieved when the nature; of the scattering objects affect

band structure formation as little as possible (i.e., absence of single object resonances),

for instance, when the size of the scatterers is small with respect to the typical distances

of the structure.

Further typical for QMC's is the possible (co-)existence of extended and localized (or

confined) as we'll as critically localized modes. While in periodic structures all modes arc;

extended unless disorder is introduced, perfect quasiperiodic order seems to get along

well with localization. This fact is usually explained by the conflict of aperiodieity, which

drives for locali/ation, and self-similarity, which drives for extended wave functions.d'19 An

intermediate, weaker form of localization is reflected m the usually power law decaying,

critical wave functions. While this connection is intuitive, the nature of the wave functions

are strictly determined by the type of the spectrum and critical waves are so intrinsic to

systems with singular continuous spectra (see Kohomoto and SutherlanddG8 and references

therein).

Similarly to all this, high-symmetry patches (clusters) with a high local scatterer

density are generally assigned the prominent role to act as centers hosting localized reso¬

nance modes (coupled single object resonances). Due to the repetitive properties of some

quasiperiodic structures, such clusters will occur everywhere in the structure, again and

again. For instance, in case of the regular Penrose tiling any patch with diameter d will be

found again within a distance of 2d. Overlapping of wave functions localized at adjacent

and not too distant clusters then allows exchange of energy and therewith propagation.

Consequently, if these clusters are not distributed sparsely (e.g., singular tilings with one

high-symmety cluster in the center) the modes are trapped. This has been studied for

QPTC-8, -10, and -12 by Wang and co-workers.d-159

11

Defects and system sizes

Due to the larger number of local environments in quasiperiodic as compared to periodic-

structures many different point defects can be formed and therewith many different defect

states. Many of the strange properties of quasiperiodic and other deterministic aperiodic

structures only develop in infinite system sizes. Even the plain Fourier spectrum of a

system of experimentally or computationally realizable sizes may look rather unspectac¬

ular. In QMC's, the number of scattering objects was as large; as a few hundreds to

thousand. Whether or not much larger structures would be closer to ideal case depends

on their experimentally achievable perfection (degree of long-range order). It determines

its structural correlation or coherence length, which should best be at least of the order

of the coherence length of the probing waves used.

2.1.2 ID QPTC's and QPNC's

The effect that the special types of order have on the propagation of waves in MC's are

illustrated by ID substitutional sequences, which represent the different ordering types.

The quasiperiodic Fibonacci sequence is thereby compared to the critically periodic Thue-

Morse and the almost periodic period doubling sequence. The ID substitutional sequences

used are words defined by a finite alphabet (A, B) a seed and a substitution rule a, which

can be applied to a word. Multiple application of a. wn — <Jn{Ä) leads to longer and

longer sequences. A overview of the properties of QPTC's based on such sequences is

given by Albuquerque and Cottam.'16

All PNC's treated in this overview consist of arrays of thin epoxy sheets in water

separated by two distances A and B which are chosen as 1 and r. The thickness, d, of

the sheets is oik1 tenth of A.

Fibonacci sequence (FS)

The Fibonacci sequence is based on the two-letter alphabet (J4, B) and the substitution

rule a (A) = B7 o(B) — BA. The substitution rule can be written in the form

with the substitution matrix S =. (2.2)

The eigenvalues of the substitution matrix result as solutions of the equation det\S —

AI| — 0 with the eigenvalue A and the unit matrix I. The evaluation of the determinant

leads to the characteristic polynomial A2 — A — 1 = 0. Tin; roots are the eigenvalues

Ai = —-— = t and A2 = —-— = —. (2.3)I It

12

If the characteristic polynomial has integer coefficients then the eigenvalue's aie always

algebraic numbers (Pisot numbers) and the sequence is quasiperiodic. One eigenvalue

is always larger than one while the modulus of its conjugate is smaller than one,d83

Ai > 1,|A2| < 1.

The structure grows as A -> AB -v ABA -> ABAAB -> ABAABABA etc.. The

length of the sequence at the iteration n is Fn+1 + Fn. The ration of the occurrence1 of

the two segments is Fn+1/Fn and conveiges towards r. If we assign intervals of lengths 1

and r to A and B then the resulting ID structure s(r) is invariant under scaling by r",

s(rr) = s(r). And so is the Fourier spectrum. The Fouiier module of the FS is of rank 2

and its diffraction pattern is a pure point spectrum (i.e., Bragg leflections only). The FS

has a periodic aveiage structuie with a period «pas = (3 — r)A

Only one year after the discovery of quasicrystals in 1984d 122 Merlin et al.d 8Tinves¬

tigated a FS based GaAs/AlAs heterostructure by Raman scattering. The theoretical

analysis of this system followed in 1987.di49 They found gaps in the density of states

of longitudinal-acoustic phonons propagating perpendicular to the multilayer structure.

Thus, in principle the fiist gap in a ID QPNC was found already two years befoie the

invention of band gap materials.

The early works on FS-QMC's clearly focused on the fundamental aspects and impli¬

cations of quasiperiodicity. First of all, it was demonstrated that band gaps can form in

non-periodic structures [see also Fig. 2.1(c)]. Kohomoto et al.d>db* further confirmed

the self-similarity and the critical nature of the wave functions in FS-QMC's in analogy to

theii previous woiiva on quasipeiiodic election systems (see also Hattori et o/.d48). An il¬

lustrative way to directly visualize the different propagation of waves in quasiperiodic and

periodic structures is to combine them in a hybrid structure. Montalban et a/.d9i have

shown how some of the waves localize in the quasiperiodic section of the structure. They

also pointed out that the bandgaps of FS-QPTC are very robust against imperfections

occurring in experimental realizations. The size-dependence of the transmission spectra

and band structure was explored by Kaliteevski and co-workers.d 62They showed, that

despite the fact that the main gaps of the band structure become cleaily visible already

for small generations of the sequence (see also Fig. 2.8), the exponential decay of the

waves inside these gaps is much weaker than in periodic structures. Additionally, in this

work a method to solving the wave equation by moans of an expansion of the fields and

the dielectric constant distribution in terms of stiong Fouriei coefficients in the diffrac¬

tion pattern of the FS. This connection appears veiy reasonable when the tiansmission

curve of FS-QPNC's are compared to the Fourier transform of its impedance distribu¬

tion (see Fig. 2.1). For every strong Bragg reflection, there is a conesponding gap And

also the fine stiucture of the tiansmission spectrum are correlated with diffraction peaks.

Furthermore1, the diffraction pattern of the FS is not only self-similar but also invariant

under scaling with factor r. The same applies to the transmission curve [Fig. 2.1(c)].

13

S (a)

|(b)

fa„„ /c

Figure; 2.1: In (a) the impedance variation of the w8 Fibonacci QPNC is shown (55

sheets) and its Fourier transform in (b). The thickness ratio of B and A blocks is 1/r.The transmission spectrum and superposed its r-scaled equivalent is shown in (c). The

inset in (c) indicates the self-similarity of the gap positions.

In the light of this analogy the fact that no connected band of strong transmission can

survive if larger and larger FS with more and more diffraction peaks are considered seems

only conclusive. The analogy of Fourier and transmission spectrum may not be very well

visible in certain realizations of FS-QMC's. The influence of the nature of the scattering

unit is best reduced by using thin interface films separating blocks of a single material

but with two different thicknesses which arc; both large with respect to the interface layer

thickness. Investigations of FS-QPTC's did not only aim for answers to fundamental

physical issues, they also proofed, that quasiperiodic structures can be well used for ap¬

plications of MC's. It was already indicated by Merlin et al.d 87 that FS-QMC's may have;

gaps which an; independent of the angle of incidence in an extended range. Indeed, a

proper design was found for an omnidirectional reflector for electromagnetic waves in the

infrared region. Lusk and Placido084 have measured an angle-independent reflectivity of

99.5 % in a FS-QPTC consisting of only 13 layers of Si02 and Si. The possibilities to

create high transmission modes in ranges of the band gaps also in aperiodic layered struc¬

tures was demonstrated by Peng and co-workers.di03 As a prerequisite for the existence

of such modes they mention the mirror symmetry of a stacking sequence. In Fibonacci

sequences mirror symmetry can be easily obtained by removal of the first two letters of

the sequence.d r'2

Typical features of a Fibonacci QPNC are shown in Fig. 2.1. The Fourier transform

of the finite ID w& structure consisting of 55 A and B blocks features already a large

number of Bragg peaks. Each of these peaks clearly induces a gap in the transmission

spectrum. The; depth as we1!! as the width of the gaps are clearly following the intensity

14

of the Bragg peaks [Fig. 2.1(c)J. The main gaps develop already in very short sequences

(i.e., the second or third generation) and do not change significantly for larger sequences.

In the ranges of the bands in between though, an increasing degree of fine structure of

gaps and peaks evolves. The positions of gaps is invariant under a scaling operation with

factor t just as is the Fourier spectrum. Also the transmission bands (which can be

determined in spectra of short sequences) show a scaling symmetry with respect to their

center as well as self similarity at any frequency. For an eighth generation sequence the

first gaps appear at foyAS/c ~ 0.2, well below the first gap induced by the Bragg peak of

the reciprocal average structure.

Thue-Morse sequence (TMS)

The (Prouhet-)Thue-Morse sequence is based on the two-letter alphabet (A,B) and the

substitution rule o~(A) = AB, o~(B) — BA. The substitution rule can be written in the

form

with the substitution matrix S —

The eigenvalues of the substitution matrix, A: = 2 and A2 — 0 can be obtained

from the characteristic polynomial A2 — 2A = 0. Despite the fact, that they are Pisot

numbers, the Fourier spectrum of the Thue-Morse sequence is singular continuous and

the sequence! is not quasiperiodic. At first glance the sequence appears to be even more

periodic than the FS. Partitioning of the sequence into AB and BA blocks results in a

periodic substructure with period A + B. Only due to the special order of the sequence

the Bragg peaks associated with the reciprocal average structure vanish for large enough

sequences as do all other Bragg peaks.d"51 The sequence grows as j4 -+ AB —» ABBA

—> ABBABAAB etc.. The length of the sequence at the iteration n is 2". The frequencies

of the letters A, B in the sequence are equal.

The facts that the Fourier spectrum of this sequence lacks Bragg peaks and that the

formation of band gaps is closely connected to Bragg scattering have spurred a surprisingly

large1 number of studies on PTC's with TMS structure1. In a very detailed theoretical

study erf the optical transmissiem as functiem e>f e>f the layer number Riklund and Severin

have shown that the transmission spectrum differs considerably from those of FS based

QPTC's in that there are certain ranges which remain almost unfragmented also for very

large numbers of layers/1110 In FS-QMC's, on the other hand, all bands are split by large

numbers of narrow gaps for sufficiently large systems [e,e)mparei Figs. 2.1(c) and 2.2(c)]. In

the; same frequene;y ranges there are almost no Bragg peaks in the Fourier spectrum. This

could be an onset of vanishing of Bragg peaks as is anticipated for the infinite sequence

(2.4)

15

S (a)

a„, /r

(b)

kUAhLk*.... ..^UjMJtjLLuJL^^JiiL

faoas /c

Figure 2.2: In (a) the structure of the w6 Thue-Morse PNC is shown in (64 sheets) and its

Fourier transform in (b). The transmission spectrum (c) again follows directly the Fourier

spectrum. Interesting to note are the ranges of strong transmission around /apAs/<" ~ 1

and 2.25. The inset in (c) shows the scaling of the spectrum with the factor 3.

(see also the section on random sequences Sec. 2.1.2). Closer to the band edges the

fragmentation of the transmission spectrum is similar in the FS and TMS MC's.

A clear distinction of classical gaps and gaps with fractal nature was suggestedd5&

in connection with the periodic average structure. A partitioning of the sequence in

groups of n members creates either a unique type of interfaces (e.g., always A\B or always

B\A) between subsequent parts (a PAS, creating classical gaps) or yield a complicated

sequence of different interface types (an aperiodic structure creating the fractal type gaps).

Strong Bragg reflections remain present in the TMS Fourier spectrum up to lengths of

several thousand units. The really special spectral property of this sequence (absence

of Bragg peaks) can thus rarely be observed in MC's, however, properties which are

clearly different from those of systems with Fibonacci structure can be seen. From a

point of view of applications it is interesting to note that TMS-PTC's with sufficiently

high impedance; contrast can have omnidirectional band gaps for electromagnetic waves

as has been demonstrated experimentally.d'24 For the low impedance contrast, such as

achievable with high- and low-porous Si hcterostructures, no omnidirectional band gaps

were found.d-t)2

Period-doubling sequence (PDS)

The period doubling sequence (PDS) is based on the two-letter alphabet (^4, B) and the

substitution rule a(A) = AB, cr(B) — AA. The substitution rule can be written in the

16

(a)

=

r/ani60

/r

" "Y YMV)

fa„„ /c

Figure 2.3: In (a) the impedance distribution of the PNC based on a Wq period dou¬

bling sequence is shown (64 sheets) and its Fourier transform in (b). The transmission

spectrum is strongly correlated with the Fourier spectrum. Interestingly, thcie are broad

bands, which are almost unfragmented and this despite the lesser degree of periodicity as

compared to the FS.

form

a

with the substitution matrix S (2.5)

The eigenvalues of the substitution matrix can be obtained from the characteristic

polynomial A2 — A — 2 = 0. The solutions aie Ai — 2 and A2 — — 1 and therewith no Pisot

numbers. The Fourier module is of infinite rank, the Fouiier spectrum is atomic (Bragg

peaks only). This means that the sequence is almost periodic. Therewith this sequence

can be categorized somewhere between the FS and the TMS The sequence grows as

A -> AB -> ABAA -> ABAAABAB etc. The length of the sequence at the iteration n

is 2n. The occurrences of the letters A, B in the sequence are as 2 to 1.

There are only a few studies on the spectral properties of the PDS-MC'sd 6'd 10, d 1 ^ in

which according MC's show properties similai to those of TMS systems. The character¬

istics of a PDS based phononic crystal an1 shown in Fig. 2.3. Cleaily this PNC transmits

much more similai to a periodic PNC than the quasiperiodic Fibonacci system does.

There are broad ranges of bands with almost full transmission. The fractal fragmentation

is restricted to certain bands (i.e., around faPAi,/c ~ 1.5 to 2).

17

Other aperiodic sequences of interest

The Rudin-Shapiro sequence bases on a four letter alphabet (A, B, C, D) and the sub¬

stitution rule a(A) = AB, a(B) = AC, a{C) = DB and a{D) = DC Alternatively a

reduced alphabet A, B —> 0 and C, D — 1 can be used. The Fourier spectrum of this

sequence1 is absolute continuous and therefore the sequence was investigated in form of

MC's several times.d 10, d ^ d i55Multilayer structures based on a Cantor set distribution

were investigated by Laviinenko and co-workers.d 7G The self-similarity of the spectral

properties weie demonstrated. In a similar work by Monsoriu et a/.d90 a strong impact

of a PAS was observed.

Modulated structures (MS)

The interesting aspect of modulated structures is that additionally to the fundamental

period, a a second scale is introduced by the period of the modulation, Amod- Modulations

can be introduced either by a periodically varying displacement of scatterers from the

basic structure or by an additional modulation of the materials properties in a MC.

Variation of the two length scales allows to significantly change the properties of a MC.

Also aperiodic sequences can be realized, of course, if the1 ratio of the two periods is

inational [incommensurately modulated structures (IMS)J. While the Fouriei spectrum is

pure point for commensurate modulations (CMS), according to Piange et al.,d lü4 IMS aie

simple prototypes for structures with singular continuous spectra (all depending on a =

^mod/tt)- The Fourier module of a dD modulated structure is of rank (d + m) with m tb°

number of wave vectors needed to describe the modulation. The Fibonacci sequence can

equally well be described as quasiperiodic or as incommensurately modulated structure.

The Fourier spectrum of a periodic MC with a sinusoidally modulated impedance of the

scattering objects [see Fig. 2.4(a)! ('an have just one pair of satellite peaks accompanying

each reflection of the basic structure. There is also such a satellite peak to the reciprocal

origin and this satellite can be very intense.

The transmission spectra foi an IMS and a CMS are compared in Fig. 2.4. The

spectra are very similar at low frequencies. The satellite peaks of each Bragg reflection

introduces satellite bands to either side of the main gaps for both structures. Towards

higher frequencies moie of the aperiodic nature of the IMS come into play. The higher

bands do not reach full transmission anymore. With regard to applications, the most

interesting question concerns the satellite to the zeio reflection, which is very strong in

the Fourier spectrum. Unfortunately, the resulting attenuation peaks in the transmission

spectrum are only very weak (about 10% for the IMS and 5% for the CMS). The same

also applies to quasiperiodic MC's because also in their Fourier spectra there are Bragg

peaks in close proximity of the origin. Modulated structures are certainly a appropriate

means to study such low frequency dips.

18

(a)

20r/a„

It

wMMT-~-»"Vî

If1kn(c)

1 WJw ,g -40

Figure 2.4: In (a) the impedance distribution of a periodic PNC's with additional

impedance modulations by frmod = 27r/r4apAs (black) and &mod — 27r/5apAs it> shown

(60 sheets). In the Fourier spectrum, each Bragg reflection is accompanied by satellites

to either side (b). The stiong satellite to the zeio peak is especially inteiesting because it

appears at very low frequencies. In the transmission spectrum there aie gaps according

to the Bragg peaks in the Fourier spectrum (c). The gaps caused by the first satellite

reflections, though, are only weak.

Random array-

While it may be easy to imagine periodic and disordered structures the aperiodic sequences

described in the previous sections are not intuitive. In order to clearly mark the extreme

cases the example of a random substitutional sequence shall be added. The sequence is

formed by random arrangement of two blocks A and D of thicknesses 1 and r. For large

sequences the Fourier spectrum becomes absolute continuous.

The characteristics of a PNC based on a random arrangement of blocks A and B,

realized two different distances in an array of thin epoxy sheets in water, are shown in

Fig. 2.5. The fourier spectrum (b) is, of course, not continuous at this size of a system, but

compared to other structures, the Bragg peaks are considerably smaller with respect to

the background intensities. The transmission spectrum features well defined band gaps.

The bands, on the other hand, show reduced transmission especially toward the band

edges. It is interesting to note, that some of the bands are much more affected by the

disorder. The bands that showed strong transmission also in the PNC's based on the

TMS and the PDS are strong also in this system and the band below /apAs/c ~ 2 shows

considerably reduced transmission in all these cases.

19

Figure 2.5: In (a) the structure of the PNC based on a random substitution sequence is

shown (60 sheets). The Fourier spectrum still features some Biagg-like intensities, (c)The transmittance spectrum (averaged over a small ensemble of sequences) follows the

Fouiier spectium. The gaps are not so much affected by the lack of order, but the bands

are. As in the PNC based on the PDS the band above the second stronger gap is most

affected by the disorder.

2.1.3 2D and 3D QPTC's and QPNC's

2D quasiperiodic PNC's and PTC's

What substitutional sequences are to ID aperiodic stiuctures are tilings to 2D aperiodic

structures. These arrangements of two or more unit blocks of usually rhombic shapes are

thoroughly studied and a vast majority of 2D quasipeiiodic MC's just uses tiling vertices as

positions of the scattering objects. Only very few investigations address the different dec¬

orations of tilings. Tilings are also used in solid state physics because they provide good

model structures for real quasicrystals. Consequently, many studies were already per¬

formed before on the propagation of electrons or general waves in such structures.d 67> d 109

In the first experiment on QPNC's performed by He and Maynardd 49 in 1989 acoustic-

tuning forks were positioned on the vertices of a Penrose tiling and coupled by thin wire's.

After this somewhat exotic start, 2D quasiperiodic structures were investigated mainly

in photonics. Chan et al.6 19 started with an QPTC-8. QPTC-10 and -12 were theoret¬

ically but also experimentally studied in 2000 by Jin et a/./157 Zoorob et al.d172 and

Kaliteevski and co-workers.d 61 In 2002 Lai et o/.d 72reported on the fiist QPNC-12 based,

in analogy to QPTC's, on wave scattering. Othei symmetries followed in a later stage.

In general investigations of 2D quasiperiodic QMC's are based on the two fundamen¬

tal gap formation processes described in Sec. 1.1.2, namely resonance basedd 60, d in and

Bragg scattering based gap formation.d6t QPNC's characteristic foi each mechanism are

studied in detail in Sees. 2.2, 2.3, and Sec. 2.5 and shortly illustrated in Fig. 2.6. In the

following the two approaches are illustrated for an octagonal QPNC and in the remainder

20

of the section more exotic aperiodic structures are discussed.

Other 2D aperiodic structures

A very simple approach to isotropic band gap materials was suggested by Horiuchi and

co-workers.d-52 They have investigated curve-linear PTC's (CPTC's) which consist of con¬

centric rings of equally spaced cylinders. They can be arbitrarily isotropic and the local

environment surrounding the individual scatterers can be quite simple square or trian¬

gular structures. Nevertheless, such circular structures do no longer have a pure point

Fourier spectrum. But a discrete spectrum is of importance for the existence of Bragg

scattering induced band gaps (see Sees, 1.1.2 and 2.3). Nevertheless, dips in the transmis¬

sion spectrum were observed indicating an omnidirectional band gap (see also Zarbakhsh

et o/.d167).As a veritable alternative to dodecagonal structures David et o/.d'16 have proposed

PTC's based on Archimedean tilings. These structures are related to dodecagonal square

triangle tilings in that they consist of the same building blocks and also have similar local

arrangements in common. As a type1 of approximant, they offer a way to achieve isotropic

band gaps (due to their local 12-fold symmetry) in less complex, periodic structures.

Also in 2D certain fractal structures were investigated by Li and co-workers.<m A

photonic crystal based on a Sirpinski structure was arranged using coaxial cables. Their

findings illustrate how the basic structure is reflected in the transmission spectra and

also how the remaining bands are more and more fragmented with increasing size of the

system. The fabrication of real MC's based on such structures is difficult due to the

very different distances that occur (i.e., these structure do not generally obey the Delone

condition). On the other hand, this is also the main motivation for constructing such

MC's. The different inherent scales can lead do gaps over a broad frequency range. This

was demonstrated also by Sheng and Chan/1125

For all the structures mentioned in this section mainly MC's with resonance-based

gaps are interesting because this mechanism is less demanding for a structure. Thereby

the question about the transmission properties of MC's with random structures naturally

arises. Random structures should, on average, be fully isotropic for infinite systems. In

principle it should be possible to create MC's, which can be described by a tight-binding

Hamiltonian, the transfer integrals of which are very small with respect to its on-site

energies (i.e., very weakly coupled resonances of the single scattering objects). The exact

structure of the system becomes increasingly unimportant the smaller this ratio is. The

factor which is directly related to the coupling strength is the volume fraction of the

scatterers in a MC. For diluted systems the formation of narrow gaps at the resonance

frequencies of the scatterers can be expected. More densely packed systems with smaller

intcr-scatterer distances should stronger respond to the lack of order and localization of

waves can be expected to gradually blurring the clear band structure in a transmission

21

(a) 20

15

10

5

coCO

e_O D

C C

aj g-4-* '-I—'CO ÜO CDCO CO

* * • • • •

t 9—9 * * »

5 10 15 20

r/apas apas ' r

CQ

c

g'cow

"ECOc

cc

0

-20

-40

-600.1

i 1 1

0.2

f r / ccy|

0.3

f apas' c

Figure 2.6: Illustration of the two gap formation mechanisms for two different QPNC-8's,

one based on soft polymeric rods in water (filling fraction 0.17) and the other consisting

of steel rods in water (filling fraction 0.27), both according to the same structure (a).Inthe scattering cross section of the soft polymeric rods (b) strong resonance states can

be observed. These resonances directly induce band gaps in the respective QPNC-8

(c). The steel rods, on the other hand, do not have such resonances in the frequency

range of interest. The structure solely determines the frequency ranges of the band gaps.

Since these evolve from Bragg scattering, the mid gap frequencies of the first gap (f) is

indicated by a strong Bragg peak in the Fourier transform of the structure; (e). This Bragg

peak further defines a periodic average structure from which the quasiperiodic structure

deviates astonishingly little (d). The maximal displacements are bound by octagons on

every node of the average structure. For some of these octagons there is no corresponding

tiling vertex (see Sec. 2.3).

22

spectrum (see Rockstuhl et al.dAU and Kaliteevski et al.dAU). In ordered systems, on

the contrary, stronger coupling rather leads to more distrinct band gaps. On the other

hand, if the filling fraction is reduced, the frequency of the first Bragg gap, fBa ~ c/2a0,

decreases too and the structure; becomes more important again due to the; overlap of the

two different scattering regimes. Thus, intermediate filling fractions appear to be most

suitable for random MC's.

3D QPNC's and QPTC's

Studies on MC's with 3D quasiperiodic structures are limited to different realizations of

the icosahedral structure. This structure is discussed in detail in Sec. 2.4.

2.1.4 Applications

While some of the applications based on QMC's can equally be produced from periodic

MC's, there are several aspects inherent to quasiperiodic structures, which make them

especially suitable. Especially the creation of wave guides can profit from the isotropy

and assumed defect insensitivity of QMC's. Waveguiding properties of QPTC-8 were

studied by Cheng and co-workers/120 The influence of the quasiperiodic structure in simple

linear waveguiding systems is difficult to be localized. A higher frequency selectivity was

proposed1' 20 but in general QPTC's mainly provide; omnidirectional band gaps and are

therefore suitable hosts for waveguides with bends/138 Waveguides with up to 180° bends

were shown to be realizable in CPTC's by Zarbakhsh and co-workers.dlf)T A special wave

guide using the coupled resonance states of high symmetry clusters in a QPTC-12 was

presented by Wang and co-workers/1-160 This system, a non-rational approximant of a

QPTC-12, is special in that it does not really contain defects.

More of the quasiperiodic structure can be exploited for the creation of cavities. The

many different vertex coordinations as well as the existence of round cluster motifs, which

can be removed to form resonators, provide a high flexibility for tuning of the frequencies

of corresponding defect modes. d,2° Resonators between adjacent waveguides in QPTC-8

were used to create add/drop filters as was impressively shown by Romero-Vivas and

co-workers.d113

Lasing sytems based on quasiperiodic PTC's wen1 investigated by Notomi et al.d,m

for penrose structure and for dodecagonal QPTC's by Nozaki and co-workers.'1-97 Non¬

linear effects like higher harmonics generation were analyzed for instance in studies by

Zhu et a/.d-m for FS-QPTC by Bratfalean el a/.d15 for penrose structures and for octag¬

onal QPTC's by Ma and co-workers.d-85 An increase in intensity of light extracted from

PTC-based light-emitting diodes was shown to result from changing the PTC structure1

from triangular periodic to dodecagonal quasiperiodic/1169 Superlenses for acoustic waves

exploiting the negative refraction index of QPNC-8, -10, and -12 were analysed by Feng

23

et al.d 35 and by Zhang*

2.1.5 Fabrication of QPTC's and QPNC's

There are many relatively simple methods to create a PNC or a PTC. Arrays of holes can

be mechanically drilled into a flat substrate, surface gratings can be machined or also fibers

or rods can be piled and mechanically connected. The same structuring possibilities are

also provided by etching techniques. These become more and more important the smaller

the constituents of the MC's become, namely in photonics. With these tools almost any

structure can be fabricated in reasonable perfection. However, they appear very inefficient

when thousands of building blocks are required. Techniques which exploit for instance a

naturally occurring ordered system like the inversion of opals or arrangements in colloidal

systems are far more elegant and efficient/1106

However efficient self-assembling routes are they will not likely yield a quasiperiodic

structure. But also other ways of efficient processing have been developed. Optical inter¬

ference lithography (holography) certainly has to be mentioned in this respect. Several

laser beams can be arranged in such a way that their interference pattern is very close

to a quasiperiodic structure.'1'13' *i58 The sets of lines of maximal intensity in such a pat¬

tern form an n-grid. This n-grid is equivalent to rotated periodic average structures of

the quasiperiodic structure, which arc spanned by symmetry equivalent Bragg peaks/1137

The number of laser beams required is equivalent to the order of diffraction symmetry of

the desired structure. Now, the energy accumulated at the maxima of such an interfer¬

ence pattern can be used to process (polymerize) a photo sensitive material. Thereby a

quasiperiodic structure can be obtained in a few seconds time (an overview is given, for

instance, by Escuti and Crawford*33). The quality of the final product can then again

be checked by laser diffraction. Instead of using multiple lasers Yang et al. have sug¬

gested that the interference pattern can also result from diffraction of a single beam on a

mask*148'*105

Interference patterns of multiple lasers can also be used to manipulate diluted col¬

loidal suspensions. Roichman and Gricr*112 have managed to arrange silica spheres on a

quasiperiodic tiling this way. A similarly versatile technique is laser-focused atomic de¬

position (Jurdik et alS1). Atoms, ablated from metallic targets, are guided by standing

waves of multiple laser beams. Thereby, the beam of atoms is focused in the nodes of the

standing waves.

A less elegant but far more flexible way to produce such grids of photoresists is given

by the direct laser writing procedure'1"77 (a collection of articles on this method can be

found in a recent MRS Bulletin*93). In this approach a single low power laser beam is

scanning the whole volume of the sample and by focusing of the beam onto certain spots

the polymerization of the resist is initiated. The resulting grids can be used as templates

for immersion processing (infiltration with a different material and subsequent chemical

24

resolving of the original mask).

Summary

The quite extensive work on quasiperiodic and other aperiodic heterostructures covers a

large variety of different topics. It is fascinating how the mathematics of fractal structures

or the theory of diffraction spectra find their way into the design and practical construction

of an exciting class of materials. The exotic aspect of multifractality can find a direct

implication on the potential of a ID MC (for which transmission and Fourier spectrum

can be very directly correlated) as broad band wave shield with many spectral gaps. The

arbitrarily high rotational symmetry and the still point diffractive spectra of 2D and 3D

quasiperiodic structures can be practically exploited to manufacture isotropic band gap

materials, which are perfectly well suitable to host wave guides or cavities. With the

future capabilities to grow largei and larger sections of QMC's the true nature of their

properties may be even better appreciated. And with the massive promotion they enjoy

currently, the reluctance to use quasiperiodic structures for further technical research on

the host of applications promised could also lessen.

25

2.2 QPNC's and single scatterer resonance states (Ar¬ticle 4)

It has been mentioned before (see Sec. 1.1.2), that strong resonance; states of the single

scattering objects can cause the formation of band gaps in the acoustic dispersion relation

of PNC's. In this article, this is demonstrated for QPNC's. More detailed investigations

of the fields eradiated from rods at resonance frequencies help to understand how this

exactly happens. Thus, the focus is on the properties of the single scattering objects.

D. Sutter and W. Sterner, Phys. Rev. B 75, 134303 (2007).

26

Prediction of band gaps in phononic quasicrystals based on

single-rod resonances

Daniel Sutter-Widmer* and Walter Steurer

Laboratory of Crystallography, Department of Materials,

ETH Zurich, 8093 Zurich, Switzerland

(Dated: March 5, 2007)

Abstract

Band-gap formation in two-dimensional quasipcriodic polymer/water hcterostructures (with 4-

to 14-fold Patterson symmetry in this study) is governed by strong acoustic resonances of the

sound-soft single scatterers. Already with an eightfold-symmetric structure the first band gap is

very isotropic. For isotropy of the higher gaps higher-symmetric structures are required. However,

this can also be achieved by a smart tuning of the properties of the scatterers. Their symmetry

(and therewith the symmetries of the scattered fields) has to better match the symmetry of a given

structure. Polygon- and star-shaped prisms on quasipcriodic structures can yield smoother and

more isotropic gaps in transmission spectra.

PACS numbers: 46.40.Cd, 01.44.Br, 43.35.+d

* Electronic address: [email protected]

27

INTRODUCTION

The study of classical wave propagation in periodic heterostructures, i.e., photonic

(PTC's) and phonemic crystals (PNC's), started almost 20 years ago [1]. Since then, the

promising applications such as optical computers and devices have spurred an almost expo¬

nential growth of the number of publications on PTC's [2] Far less work has been devoted

to PNC's. For these, potential applications are expected in noise control and ultrasonic tech¬

nology, for instance The similarity of PTC's and PNC's allows, to some extent, a knowledge

transfer and increases the impact of discoveries in each field. The fascinating type of com¬

posite materials can be described as one-, two- (2D), oi three-dimensional meta crystals

built of objects which scatter electromagnetic or elastic (acoustic) waves if the wavelength

is on the scale of the lattice period (for a comprehensive review, see Ref. 3).

The existence of omnidirectional band gaps, which is important for most applications, is

strongly favored by high symmetries of the heterostructures. The rotational symmetry of

periodic structures is limited to sixfold. For 2D quasiperiodic structures there is no upper

limit and consequently quite a few publications already report the peculiarities of quasiperi¬

odic PTC's (QPTC's) and PNC's (QPNC's) (see Rcfs. 4 8 and references therein). However,

bands and gaps m QPNC's are well defined in particular cases only (l e,in some systems

only pseudogaps were found [9] similar to the electronic pseudogaps of real quasicrystals)

and their formation and structure is not yet thoroughly understood In the following, we

present a study of the scattering properties of single rods and show how this information

supports the understanding of the formation and the optimization of band gaps in QPNC's

The transmission spectra for a square lattice PNC as well as QPNC's with 8-, 10-, 12-, and

14-fold Patterson symmetry (see Fig. 1) were calculated by a finite difference approximation

in the time domain (FDTD) [10]. For the scattering cross-section calculations of cylindri¬

cal rods we have used a multipole-expansion method [11] and for all other rods the FDTD

method.

I. SYSTEMS OF CIRCULAR CYLINDRICAL RODS

The type of scattering in PNC's has been known to be of prime importance ever since the

first PNC's were created. It can be adjusted by the impedance contrast of the constituent

28

FIG. 1: Quasiperiodic structures with 8-, 10-, 12-, and 14-fold Patterson symmetry considered in

this study. The arrows in each pattern designate the independent high-symmetry directions.

phases as well as by the volume fraction of the scattering objects. Especially in systems with

hard contrasts and sparse scattcrcr distributions, the mechanism for band-gap formation is

based on Bragg scattering. Strong Bragg peaks in the Fourier spectrum of the underlying

structures directly indicate the possible frequency ranges of the band gaps [8, 9]. On the

other hand, in soft-contrast systems with sufficiently high filling fraction, the resonance

modes of the scattering objects can play a very dominant role in determining the frequency

ranges of band gaps [the approach was used early for PNC's (Ref. 12) and recently also

applied to QPTC's (Ref. 13)]. The resonance frequencies are independent of the structure,

instead they scale with the speed of sound in the material of the scatterers and inversely

with their size. The coupling of such resonance states in a QPNC spreads these states to

form a band. The interaction of this band with the continuum band of the effective medium

produces a band gap due to hybridization (for a very clear description of this mechanism see

Ref. 14). The correlation of resonance frequencies and gap positions is shown in a comparison

of PNC's and QPNC's of 4-, 8-, 10-, 12-, and 14-fold Patterson symmetry (Fig. 2). The

heterostructurcs consist of polymeric rods (^=1800 m/s, vs=800 m/s, p = 1.14 kg/m3)

in watei at filling fractions of 0.17. Samples of about the same thickness in direction of

transmission were set up with 357, 361, 365, and 355 rods for the QPNC's with 8-, 10-, 12-,

and 14-fold Patterson symmetry, respectively. Similar to what has been found by Rockstuhl

et al [13]. for photonic systems, the band gaps occur at frequencies close to those of the

29

resonance states in the scattering cross sections of a single rod. Nevertheless, in these

(Q)PNC's the arrangement of the rods does play a crucial role. For the periodic square

lattice PNC the first band gap is shifted by almost as much as its width if the direction of

transmission is changed. A very bad overlap results. This overlap is clearly getting better

with an increasing degree of rotational symmetry of the arrangement of the scatterers. While

for the 8-fold structure mainly the first gap is absolute, for the 12-fold structure all gaps are

perfectly isotropic. The increasing symmetry of the structures also leads to broadened gaps

with less sharp edges (i.e., spikes associated with localized modes appear). This effect can

also be seen as due to more inhomogeneous nearest-neighbor-distance distributions of the

highly symmetric structures. The shorter distances tend to broaden the gap and the wider

spacings to close it. In the square structure all rods have the same coordination and thus the

overlap of their scattered field lobes with those scattered from neighboring rods is equal (i.e.,

equal transfer parameters). Thus, for the formation of isotropic and sharply bound band

gaps a structure with high Patterson symmetry and only few different vertex coordinations

seems most promising (i.e., not a random arrangement). Quasiperiodic structures optimally

combine this.

In order to predict the isotropy of a band gap in (Q)PNC's, the scattered wave field \&s

can be analyzed for the resonance, which induces the gap

Mr, 0) = e1 Y, cm{u)Jm{kr)Cos{m9), (1)

with Jm being Bessel functions of the first kind and cm the coefficients obtained from eval¬

uation of the boundary condition at the cylinder surface [11]. The index m of the strongest

coefficients in the spectrum of the expansion in cylindrical harmonics cm is indicated below

the resonance peaks in Fig. 2. These eigenmodes feature 2777-fold rotational symmetry and in

the case of a single-valued spectrum, the scattered field predominantly adopts the symmetry

of this component.

For transmission in the two high-symmetry directions indicated in Fig. 1, the scattered

waves typically encounter nearest-neighbor rods on vertices of regular ri-sided polygons (with

one vertex in the forward direction) for even and 277-sided polygons for odd n (direction of

dark arrows in Fig. 1) or just between these neighbor vertices (bright arrows). Strong inter¬

action of scattered waves (i.e., a large overlap of the scattered field lobes) occurs most likely

when the field lobes point in the direction of the nearest-neighbor rods. This interaction

30

(a)

m

O

i'Ew

c

CD

(b)

ü

COI

</-

t,,

eo

1000

200 400 600

f [kHz]

800 1000

FIG. 2: Transmission spectra for square PNC and QPNC's with different Patterson symmetry

(a). The two curves in each section correspond to the two directions of transmission indicated

with arrows of the same line style in Fig. 1. Resonance states in the scattering cross section of a

cylinder (b).

strength spreads the bands of coupled resonance states which, by hybridization with the

continuum band, produce the band gaps and determine their widths. Omnidirectional gaps

can be expected from modes with lobes of the scattered fields covering rods in the directions

of both the vertices of the n-sided polygons as well as those in between them; this is when n

is a multiple of m (e.g., the first gap in the octagonal system). Modes of low symmetry form

31

isotropic gaps in highly symmetric structures because the broad field lobes cannot resolve

the angular fine structures of the 7?-sided polygons hosting the rods. This almost guarantees

isotropy of the first gaps in QPNC's with large n. However, optimal performance requires a

good match of structure and scatterer.

II. SYSTEMS OF POLYGONAL OR STAR-SHAPED PRISMS

Due to the dominant role the properties of single scatterers play in the band-gap for¬

mation, a more detailed examination of these seems crucial. In this section we study the

influence of modified geometrical cross sections of the rods on their scattering behavior.

The shapes analyzed here are regular n-sided polygons (with constant incircle) and a five-

pointed-star. They are interesting from many points of view. First, we have seen that the

high-symmetry resonance modes do not easily form isotropic gaps. A reduction of the sym¬

metry of the scattering object can affect the symmetry of the modes. Second, for scatterers

with lower symmetry (diffeient extensions in different directions) the resonance frequencies

should change with the direction of the incident plane wave. This variation could lead to

widened gaps in QPNC's. Third, the faces of the polygons and stars form sets of broken

planes, which could give rise to a stronger interaction of reflected wave intensity.

For the polygonal prisms, the scattering cross sections for plane waves are shown in Fig. 3.

In the frequency range of interest they are very similar for cylindrical rods and for polygons

with large n. The scattering strengths as well as the Q factors of the resonances are similar

for all shapes of rods. The scattering behavior of the octagonal prism deviates from that of

the cylindrical rod only in the orientation-dependent frequency of the fourth resonance. For

the pentagonal rod more evenly spaced resonances appear, which are almost independent

of the direction of incidence of the plane wave. The square and the triangular prisms show

clearly different spectra. As anticipated, they possess more resonances at low frequencies

and these depend strongly on the direction of incidence of the plane wave. Especially for the

very first resonances, there are certain directions from which these modes cannot be excited

at all. In oblique directions though, most modes are accessible.

Now, let us have a look at how the band gaps of a QPNC's of polygonal rods look

like. Uniformly oriented pentagonal rods on the Penrose quasilattice produce the spectra

shown in Fig. 4. Compared to the Penrose QPNC with cylindrical rods (Fig. 2) this QPNC

32

o200 400 600 800 1000

f [kHz]

FIG. 3: Scattering cross sections for plane waves at polygonal prisms (shown on the right-hand

side) in different orientations (wave incident along the lines crossing the shapes).

clearly features more isotropic band gaps. Again, the gaps appear exactly at the resonance

frequencies. Due to the unsplit second peak, there are fewer gaps but instead they agree

better in their position and width for the different directions of transmission. The spectra

are also smoother than those of the cylindrical rod system around the second and third

resonances of the cylinder, which arc very close. The amplitude distribution of the scattered

field |^s| at the first resonance of the pentagonal prism is shown in Fig. 4(c). It features

well-defined fourfold symmetry.

In Fig. 5(b) the scattering cross section of a fife-pointed-star-shaped prism (incirclc 0.3

mm) is shown and compared to that of the pentagonal prism. The first resonance appears

at very low frequency. It reflects the larger maximal extension of the star and its intensity

is weak. In the arrangement of the star-shaped rods on the Penrose structure, this mode

induces only a weak attenuation peak. The second resonance frequency is almost equal to

the first one of the pentagonal rod. The scattered fields at this common resonance frequency

are similar as shown in Figs. 4(c) and 4(d) and can be further characterized by the radiation

patterns shown in Fig. 5. These patterns show the angular distribution of scattered intensity

for the far field [Fig. 5(c)] and at a distance le away from the scatter [Fig. 5(d)] (with le

being the edge length of the Penrose tiling). According to these patterns, the noncylindrical

33

FIG. 4: Band gaps in transmission spectra of the Penrose QPNC (a) and the resonance modes of

the pentagonal prisms inducing the gaps (b). The scattered fields |\I/S| for the pentagonal (c) and

the star-shaped prisms (d) at their common resonance frequency [see arrow in Fig. 5(b)].

scatterers produce slightly less sharp field lobes at both distances. Thus, slightly better

isotropies of the gaps can be expected for the Penrose QPNC with pentagonal or star-

shaped prisms as compared to those of the cylindrical system. The first two star resonances

produce highly isotropic transmission gaps in the QPNC. These gaps are again smoother

than those induced by resonances of cylindrical rods and their width is rather small. The

different widths of the coinciding gaps of the pentagonal and star systems are indicated by

the different Q factors of the corresponding resonances.

To give an example for QPNC's consisting of the more anisotropic square rods, we have

analyzed an octagonal QPNC. The orientations of the rods [see Fig. 6(c)] are chosen in such

a way that the eightfold symmetry of the structure is preserved. Corresponding transmission

spectra are compared with the different scattering cross sections of the square rod in Fig. 6.

The resonances that are accessible only in certain directions all contribute to the isotropic,

almost overlapping (and therewith broadened), first gap. Thus, anisotropic resonances can

form isotropic band gaps at lower frequencies. At higher frequencies only the isotropic modes

produce absolute band gaps. The spectra are not smoother than those of the system with

cylindrical rods but despite the reduction of symmetry of the scatterers the band gaps arc

highly isotropic.

34

FIG. 5: (a) Band gaps in transmission spectra of the Penrose QPNC consisting of star-shaped

prisms and (b) the resonance modes of single prisms inducing the gaps. Radiation patterns for

resonances indicated with an arrow in (b), for the pentagonal and the star-shaped prisms as well

as the first cylindrical resonance measured in the far field (c), and at a distance le away from the

prisms (d) (with lc being the edge length of the Penrose tiling).

FIG. 6: All resonances of the square rods contribute to the formation of band gaps in the octagonal

QPNC (a) although some of them can be excited only in certain orientations (b). The orientations

of the square rods on the tiling are shown in a quarter section of the QPNC in (c).

CONCLUSIONS

We conclude that quasiperiodic geometries are very well suited for phononic crystals

consisting of soft-contrast cylindrical rods in a liquid host. The strong resonances of such rods

govern the formation of band gaps and allow the high rotational symmetries of quasiperiodic

structures to be fully exploited to make the band gaps isotropic (in contrast to systems

without resonances [8]). In addition to the usual focus on the arrangement we have shown

35

that simpler and more isotropic transmission spectra can be obtained alternatively by using

polygonal or star-shaped rods, the scattered fields of which better match the symmetry of the

structures. The high degree of isotropy seems very promising for all types of applications of

such heterostructures, and may also encourage further analysis of new, interesting building

blocks for phononic as well as photonic crystals other than cylindrical rods.

Acknowledgments

We would like to thank B. Djafari-Rouhani, Y. Pennec, and J. 0. Vasseur for discussions

and for providing the FDTD code we adapted. We also thank Y. Psarobas and R. Sainidou

for discussions.

[1]

[2]

[3:

[4]

[5]

[6

[7]

[8]

[9

[io:

in

[12

[13

[14

E. Yablonovitch, Phys. Rev. Lett. 58, 2059 (1987).

http://ph.ys.lsu.edu/^jdowling/pbgbib.html

M. Sigalas, M. S. Kushwaha, E. N. Economou, M. Kafesaki, I. E. Psarobas, and W. Stcurer,

Z. Kristallogr. 220, 765 (2005).

Y. Lai, Z. Q. Zhang, C. H. Chan, and L. Tsang, Phys. Rev. B 74, 054305 (2006).

Y. Lai, X. Zhang, and Z. Q. Zhang, J. Appl. Phys. 91, 6191 (2002).

M. Hase, H. Miyazaki, M. Egashira, N. Shinya, K. M. Kojima, and S. I. Uchida, Phys. Rev. B

66, 214205 (2002).

D. Sutter and W. Steurer, Phys. Status Solidi C 1, 2716 (2004).

D. Sutter-Widmer, S. Deloudi and W. Steurer, Phys. Rev. B 75, 094304 (2007).

M. A. Kaliteevski, S. Brand, R. A. Abram, T. F. Krauss, R. DeLa Rue, and P. Millar,

Nanotechnology 11, 274 (2000).

B. Djafari-Rouhani, Y. Pennec, and J. 0. Vasseur (private communication).

J. J. Faran, J. Acoust. Soc. Am. 23, 405 (1951).

M. Kafesaki and E. N. Economou, Phys. Rev. B 52, 13317 (1995).

C. Rockstuhl, U. Peschd, and F. Lederer, Opt. Lett. 31, 1741 (2006).

R. Sainidou, N. Stefanou, and A. Modinos, Phys. Rev. B 66, 212301 (2002).

36

Additional notes to article 4

In the previous article the maximization of the isotropy of transmission behavior of

QPNC's is discussed. The limiting case of an infinite degree of rotational symmetry,

a PNC based on a pinwheel-tiling with a special decoration*1'12 shall be considered here.

The pinwheel tiling'1107 can be obtained by substitution of an initial triangle by five

smaller triangles (see Fig. 2.7). The particularly interesting property of this tiling is that

with every substitution step the number of different orientations of the unit tiles increases.

The rotational symmetry of the diffraction pattern grows accordingly. This makes the

structure an excellent candidate for literally isotropic band gaps. If the tiling is decorated

with one vertex per tile according to Fig. 2.7(a) the powder diffraction pattern becomes

equivalent with that of a simple square lattice.d12 As usual, tin; size of the realizable PNC's

are rather small and the Fourier spectrum of the finite structure, shown in Fig. 2.7(b),

is far from isotropic. Instead a strong reciprocal square lattice is visible as well as very

well defined octagons. The transmission spectra of a corresponding PNC, however, show

a nearly perfect direction independence [see Fig. 2.7(c)]. This clearly motivates further

studies and realizations of MC's with this structure or also other decorations (i.e., with

more homogeneous local vertex densities) of the tiling (see Parker et a/.d102).

* t * <*•

•\ * H t .• • #•! • • *

• #.'*•* •"." t •...+ X '. .

^ B• ••/ ^••••^ \

> * % • • ^ ^ • > ^ • • • • «

*,**•* • .-» •.*.* *••»

Figure 2.7: In (a) the structure of the pinwheel PNC is shown and its Fourier transform

in (b). The transmission spectra (a) are almost identical for the three different directions

indicated by arrows in (a).

37

2.3 QPNC's and Bragg scattering (Article 5)

In the overview of ID QPNC's (Sec. 2.1.2) the direct connection of the Fourier spectrum

and the transmission spectrum is demonstrated. If there aie no single rod resonances (as

discussed in the previous section) in the frequency range of interest then this connection

is valid also for 2D QPNC's. Beyond the positions of the band gaps the Fourier spectrum

provides information about the periodic average structures of a quasiperiodic pattern. In

this article periodic average structures are used to interpret the transmission spectra and

explain the occurrence 01 absence of clear band gaps.'1143

D. Sutter, S. Deloudi and W. Steuier, Phys Rev. B 75, 094304 (2007).

38

Prediction of Bragg-scattering-induced band gaps in phononic

quasicrystals

Daniel Sutter-Widmer,* Sofia Dcloudi, and Walter Steurer

Laboratory of Crystallography, Department of Materials,

ETE Zurich, 8093 Zurich, Switzerland

(Dated: January 16, 2007)

Abstract

We have studied ultrasonic wave propagation in two-dimensional quasiperiodic steel/water het-

erostructures with 8-, 10-, 12-, and 14-fold Patterson symmetry. The formation of band gaps in this

kind of quasiperiodic phononic crystal (QPNC) is mainly governed by multiple Bragg scattering.

The particular role of the periodic average structure of the QPNC in prediction and understanding

of their transmission spectra is shown. The smaller the deviations from average periodicity, the

greater is the similarity of transmission spectra of the quasiperiodic and periodic heterostructures.

Consequently, QPNC's with eightfold symmetry (QPNC-8's) show much better defined transmit-

tance bands and gaps than QPNC-10's and QPNC-14's, whose spectra resemble more those of

disordered periodic systems. Smoother transmission spectra with clear gaps can be obtained by

replacement of the cylinders of QPNC-10's by star-shaped prisms.

PACS numbers: 61.44.Br, 46.40.Cd, 43.35.+d

"Electronic address: daniel.sutterOmat .ethz.ch

39

INTRODUCTION

Research on photonic (PTC's) and phononic crystals (PNC's) started almost twenty years

ago [1]. The steadily increasing interest in this field is reflected in the nearly exponential

growth of the number of publications. The Photonic and Sonic Band-Gap Bibliography

already contains more than 8200 entries for PTC's and around 160 entries for PNC's [2].

As a consequence of the huge interest in optical computers and other devices, research in

PTC's has been much more intense and application driven. Investigations of PNC's have

been motivated mainly by applications in noise control and ultrasonic technology or by new

approaches toward fundamental physical issues (e.g., the disorder influence on band struc¬

tures). These fascinating composite materials can consist of one-, two-, or three-dimensional

arrangements of scattering objects for light or sound waves on length scales of the order of

the wavelength. Wave propagation in these materials can be efficiently controlled by the

materials employed for scattering objects and the surrounding matrix as well as their spatial

arrangement. Their dispersion relations may feature band gaps and defect states resembling

those of electrons in real crystals. PTC and PNC designs can nowadays be optimized for

specific applications, such as mirrors, waveguides, filters, etc., by the sophisticated computer

software available or by the likewise advanced manufacturing and measuring procedures [3].

The propagation of electromagnetic waves in PTC's and that of elastic or acoustic waves in

PNC's is similar in many ways and research in PNC's can profit from that in PTC's, and

vice versa [4].

The first studies on photonic crystals with quasiperiodic structures (QPTC's) were per¬

formed soon after the discovery of quasierystals [5] (the first one-dimensional QPTC was

manufactured in 1985 by Merlin et al. [6]), at first curiosity driven and then mainly motivated

by the promise of isotropic band gaps, which are required, for instance, for waveguides. One

of the advantages of quasiperiodic structures (quasilattices) is that they are not restricted

in their rotational symmetry and still maintain ideal long-range order. Besides this quest

for isotropy, there were many peculiar and exotic wave phenomena in quasiperiodic me¬

dia which, predicted by theoretical physics, awaited experimental confirmation. For the

study of one-dimensional quasiperiodic photonic systems, ideas of the fundamental work on

quasiperiodic tight-binding Hamiltonians of the time [7-9] were picked up. Layer structures

of semiconductor materials with non periodic stacking order were compared to periodic sys-

40

terns by optical transmittance investigations [10, 11]. For different phononic quasicrystals

(QPNC's) such comparisons were performed by Velasco and co-workers [12, 13]. In all these

works, clear gaps in the density of states (DOS) were reported, as was anticipated from the

analysis of one-dimensional quasiperiodic Hamiltonians. The widths of the gaps were often

found to be smaller than those of periodic systems but their number was greater. Transmis¬

sion gaps were found also in two-dimensional systems by Chan et al. [14] in photonics and by

Lai et al. [15] in phononics. QPTC's with quasiperiodicity in three dimensions were tackled

by Man and co-workers [16]. In our paper the focus is on two-dimensional quasiperiodic

structures.

With a few exceptions, almost all previous studies on such systems have been performed

on QPTC's. Soon after the first theoretical predictions of band gaps in eightfold-symmetric

photonic quasicrystals (QPTC-8's), experimental confirmation was provided for a QPTC-12

[17, 18]. On the other hand, some systems, such as QPTC-10's (Refs. 20 and 19) and QPNC-

10's [21], were analyzed which did not show clear bands and gaps in transmission spectra

but rather dips associated with pseudogaps analogous to those in the electronic DOS's of

quasicrystals. In the past years a variety of QPTC's with Patterson symmetries of order

8 [14], 10 [19, 20], 12 [15, 17], 14 [22], and 18 (Rcf. 22) have been investigated in different

material combinations and size regimes. These differences render a direct comparison of the

structures difficult.

Our previous experimental investigations [21] of PNC's and QPNC-10's (steel rods of 1

mm diameter submerged in water) by ultrasonic transmission spectroscopy encouraged us to

a systematic study of QPNC's with other symmetries. Steel/water PNC's were found to be

well capable of producing band gaps in any periodic arrangement. The first investigations

of QPNC-10's, however, did not produce any clear band features. Also the transmission

properties of QPNC-12's did not resemble those of periodic PNC's. Only QPNC-8's revealed

a stronger second band as well as two weak, narrow bands at higher frequencies (see Fig. 1).

Apart from that, all spectra consist mainly a of dense series of spikes and peaks just like

the spectra of disordered PNC's (although the structures used are perfectly ordered). A

question regarding the origin of the differences in transmission behavior, naturally arises.

This question is connected with the questions about the origin and mechanism of formation

of band gaps in QPNC's. We provide additional information for this discussion.

Wave propagation was simulated by means of finite-difference approximations in the time

41

10°

g io-1

"E

I io-2

100 1000 2000 3000

f [kHz]

FIG. 1: (Color online) Ultrasonic transmittance through QPNC-8. In red (gray) the experimental

spectrum and in black the FDTD simulation is shown. The spectrum can be separated into a low-

frequency part of effective-medium wave propagation (a), followed by the first gap (b) and a couple

of broader peaks which are gradually getting lost in the background. Above 2 MHz two bands (c),

which are associated with single-rod resonances, provide some more transmission channels.

domain (FDTD) [25], Figure 1 shows a comparison of simulated and measured transmission

spectra. The two curves agree well. Deviations of the spectra in the middle range stem from

slight differences in the setup (periodic boundary conditions perpendicular to the direction

of transmission were used in the simulations). In the remainder of this paper only simulated

curves are shown.

The patches of the tilings used in this investigation are shown in Fig. 2. The filling

fractions of the QPNC's are 0.27, 0.28, 0.23, and 0.17 for the the 8-, 10-, 12-, and 14-fold-

symmetric tiling, respectively. Different filling fractions arc enforced by a minimal rod-to-rod

distance, which must be retained for the experimental setup. To compensate for the different

filling fractions the thickness of the patches was adjusted (334, 361, 409, and 493 rods for

the four patterns). The relevant symmetry for the distribution of Bragg intensities is the

Patterson symmetry (the symmetry of the autocorrelation function). Therefore, we will

use the Patterson symmetry for the classification of tilings in the following ("n-fold" will

designate "n-fold Patterson symmetry" unless otherwise specified).

I. PERIODIC AVERAGE STRUCTURES

In order to better understand the peculiarities in the formation of band gaps, the trans¬

mission spectra of different QPNC's are compared to those of the PNCs with their periodic

42

J I I L.

FIG. 2: Quasiperiodic patterns considered in this analysis. The two arrows designate the asym¬

metric unit of the corresponding Fourier spectrum. Eight-ring clusters (structure motifs) and

the rhomtai of the scaled quasilattice, as used in Sec. IIA, are marked in the eightfold-symmetric

structure.

average structures (PAS's). For this purpose, the PAS's of 8-, 10-, 12-, and 14-fold tilings

have been derived. In Sec. IC we discuss the transition from the PAS to the quasiperiodic

structure and compare it with a transition to different disordered structures. All investigated

structures and their abbreviations are listed in Table I.

A. Crystallography of periodic average structures

The PAS of a quasiperiodic tiling is a structure, whose Fourier transform consists of

a subset of the strongest Fourier coefficients of the tiling. It can be easily obtained via

higher-dimensional description of the quasiperiodic structure [26, 27]. In this approach a d-

dimensional (dD) quasiperiodic structure is described as intersection of an nD hypercrystal

structure (with n > d), i.e., a hyperlattice decorated with hypcratoms (atomic surfaces, oc¬

cupation domains), with the dD physical (parallel) space. In Fourier space, this corresponds

to a projection of the nD weighted reciprocal hyperlattice along the (n-d)D perpendicular

space onto the dD parallel reciprocal space. The n-star of reciprocal basis vectors allows

one to index the dD diffraction pattern of the quasiperiodic structure by integers. Project¬

ing the nD hyperlattice along the perpendicular space upon parallel space gives a dense

43

TABLE I: Overview of phononic crystal (PNC) structures.

Code Description Section Figures

QPNC-n PNC with n-fold quasiperiodic structure (see Fig. 2). IB 1,6

APNC PNC with periodic average structure of a correspond- IB 6

ing QPNC.

Di PNC with displacively disordered square lattice IC 7

structure [random displacement vectors are bound

by a regular octagon as in Fig. 3(c)].

Di PNC with square lattice structure and vacancies. IC 7

Di PNC with vertices lying halfway between those of IC 7

QPNC-8 and their closest PAS nodes

D\ QPNC-8 with rods forming eight-rings replaced by IIA 8

larger rods with diameter of the ring.

set of projected vertices. The corresponding dD reciprocal space section would contain the

origin only. However, if a proper oblique projection is performed, all vertices project onto

the lattice nodes of a periodic average structure. Then, the related f/D reciprocal space

section contains a subset of the strongest Bragg Peaks. This fact is used by an alternative

approach for the derivation of the PAS, which just adopts the diffraction vectors of d strong

Bragg peaks as the reciprocal basis of the PAS. These reciprocal vectors are orthogonal to

all directions of oblique projection in real space.

In the following, we use the higher-dimensional approach for deriving the 2D PAS of the

2D eightfold tiling. Its 4D embedding space consists of two orthogonal subspaces, the 2D

parallel space and the 2D perpendicular space. The 4D hypercubic lattice is decorated with

2D atomic surfaces, which are perpendicular to parallel space and of regular octagonal shape.

The vertices of the eightfold tiling are the intersection points of these atomic surfaces with

the parallel space. Oblique projection of a unit cell of this higher-dimensional system results

in a projected lattice (the PAS) and projected atomic surfaces on its nodes. The projected

atomic surfaces on different vertices overlap for most directions of projection. Only in certain

cases they do not. These are the projections that lead to useful PAS's. For the eightfold

tiling such a PAS can be found, for instance, from a projection along (1, —1,1,0) and along

44

(a)

i

#. •. _ _ . »

i

.- .

I

'"! '

i •

• # -. i;:...

#ii

*

%

"

*

ti —....,,,. . .. —:

I.'J.VIAri.*.;..!

(c)

il&, >::oxo::::&:

FIG. 3: Fourier transform (a) of the eightfold tiling (b), with the reciprocal lattice of the selected

PAS marked with circles. The unit cell of the PAS with projected atomic surfaces is shown in (c).

Here, the projected atomic surfaces were generated by taking the modulus of the vertices of an

eightfold tiling within the PAS unit cell. With increasing size of the tiling, these forms converge

toward dense octagons, as shown in the upper left corner of (c).

(0,1, -2,1). Then, the projected atomic surface is a regular octagon as is shown in Fig. 3

and the PAS is a square lattice (SQ) with lattice parameter a0 — 2ar/(v/2 +1) with aT being

the edge length of the underlying tiling. The corresponding reciprocal square cell is spanned

by the vectors {(110Ï), (Olli)} of the 2D Fourier module

M* Hll = 5>a*|a* = a*

i=i

,hi e TL 1)

COs(27T0j)

sin(27T0j)

with z„,=4, 0i — (z — l)/8, and a* - l/2or (see the Appendix). Alternatively a rhombic PAS

(RH) can be obtained by choosing the reciprocal vectors {(110Ï), (1110)}.

The physical significance of these projected atomic surfaces (octagons) is the following.

All vertices of the tiling are lying within such an octagon situated on the vertices of the

PAS. However, not all octagons are occupied by a vertex because the vertex densities of the

quasilattice and PAS are not equal. The size of the projected atomic surface is a measure for

the displacement of the quasiperiodic pattern from its PAS (i.e. an amplitude of displacive

disorder), and the occupancy of the PAS nodes, pocc, is related to the density of vacancies,

(1 — pocc)- The smaller this amplitude is and the closer pocc is to 1, the higher is the

degree of averaged periodicity in a structure. Considering the finite QPNC's as perturbed

periodic crystals, the pattern with the smallest deviations from averaged periodicity appears

45

to be most likely the one with deep and wide gaps in the DOS, comparable to those in

PNC's. Therefore a classification of tilings according to the ratio aa of the area covered

by the projected atomic surface to the size of the PAS unit cell and also pucc is given here.

In general, a quasiperiodic tiling has infinitely many possible PAS's [27], as for instance

a*1 = a1/(y/2 + l)n, reflecting the scaling symmetry of the Fourier transform of the tiling.

The most relevant and interesting PAS's arc those with the smallest aa, with cell sizes

comparable to the tiling edge lengths (i.e., occupancy factors close to 1, because aa —» 0 for

large n), and with strong Bragg peaks. Values for aa as well as amax, the maximal, and ä,

the average, displacement of tiling vertices from the average structure are given in Tab. II

and occupancy factors can be found in Table III.

For the eightfold tiling, there is only one single atomic surface. The 10-, 12-, and 14-

fold tilings arc all generated by atomic surfaces consisting of several disconnected parts.

To illustrate the displacements of the tilings from their PAS's the modulus of all vertex

coordinates within the PAS unit cell can be calculated. For large tilings this draws nicely

the contours of the projected atomic surfaces [see Figs. 3(c), 4, and 5], The comparison

clearly reveals a homogeneous distribution of vertices in the octagon and nonhomogeneous

distributions in the projections of the atomic surfaces of the other tilings. Since all sections

of atomic surfaces have constant vertex densities (because they correspond to a projection of

an infinite str^ of the higher-dimensional lattice onto perpendicular space) a vertex density

variation in their projection must be due to overlapping projected parts (except for the 4D

atomic surfaces of the 14-fold tiling).

The tenfold tiling can be generated from a 4D hyperrhombohedral lattice decorated by five

atomic surfaces, i.e., one point at the origin of the 4D unit cell plus four regular pentagons

parallel to the perpendicular space, centered at i/bx (1,1,1,1) for i = 1,..., 4 on the the body

diagonal. It is intuitive to first project along the cell diagonal to ensure overlapping of all

pentagons and then perform a second projection to get the PAS. Choosing (4,1,1,1) as the

latter, a rhombic PAS is obtained which is decorated by the overlapping pentagons [26]. This

PAS corresponds to the vectors {(00Ï1), (01Ï0)} in the reciprocal lattice defined in Eq. (1)

with im — 4, (j)% = i/h, and a* = 2r2/5ar [with the golden ratio r = (%/5 + l)/2 — 1.618].

In the case of the 12-fold tiling, the atomic surfaces in the 6D hypercrystal structure

comprise four triangles and four nonregular hexagons and are parallel to the first perpen¬

dicular subspaee [see {v\,wi} in Eq. (A.3)] and stacked along the diagonal of the unit cell

46

TABLE II: Tilings and their PAS's parameters.

Tiling PAS lattice parameter ciq oiniaxa ah aac

eightfold SQ 2/(v/2 + l)=0.828 0.27 0.16 0.19

eightfold RH 2>/2/(V2 +1)=1.172 0.45 0.19 0.14

tenfold (3-t)/t=0.854 0.65 0.23 0.34

12-fold 2>/3/(2 + a/3)=0.928 0.128 0.073 0.14

0.104 0.052

14-fold [Fig. 5(a)] 7/{2[2 + 4cos(tt/7) + 3cob(2tt/7) + cos(3tt/7)]} = 0.455 0.27 0.44

14-fold [Fig. 5(b)] 7/{cos(tt/14)[2 + 4cos(tt/7) + 4cos(2tt/7)]} = 0.887 > a0 0.90

"Maximal displacement of the quasilattice from the periodic average structure (in fractions of o,q). For the

12-fold tiling there is one value for each of the two disconnected sections of the projected atomic surfaces.

6Average displacement of tiling vertices from the PAS.

rRatio of the area covered by the projected atomic surface to the area of the PAS unit cell.

[28, 29]. There are PAS's for which all these components overlap, e.g., the PAS correspond¬

ing to the vectors {(3402), (0232)} of the reciprocal lattice defined in Eq. (1) with im = 4,

0i — (z — 1)/12, and a* — l/3ar. The rhombic PAS spanned by the basis reciprocal to

{(2212), (Ï022)} leads to the projections shown in Fig. 4(c) in which there are regular do¬

decagons at the PAS vertices and two polygons with nine edges and threefold rotational

symmetry equally spaced on the longer cell diagonal. Due to the larger occupancy the latter

is more suitable for the comparison of phononic structures.

The 14-fold tiling can best be constructed from a 7D hypercubic lattice [30]. The atomic

surfaces are constituted by seven four-dimensional polytopes. 5D projections are necessary to

generate a PAS from higher dimensions. Directly from the 2D Fourier transform [see Eq. (l),

with im — 7, (j), — ?/7, and a* — 2/7o,] interesting PAS's can be found with combinations

of reciprocal vectors (11100ÎÏ) and its rotational symmetric equivalents. There are rhombic

PAS's with nonoverlapping projected atomic surfaces produced by {(2112202), (221Ï2Ï1)}

[see Fig. 5(a)]. A PAS with lattice parameters closer to the tiling edge length is specified by

{(lOÏIOll), (110ÎÏÏ0)}. This case is special in that the projections of the atomic surfaces

arc no longer centered on the PAS vertices [see Fig. 5(b)].

47

FIG. 4: Projections of the atomic surfaces to the unit cell of the PAS of the 10-fold (a) and the

12-fold tiling (b). In contrast to the eightfold tiling these tilings arc generated by atomic surfaces

consisting of several disconnected parts. Nonhomogcneous projections are due to overlapping of

the projections of these parts.

FIG. 5: Projected atomic surfaces for the 14-fold tiling for PAS's spanned by bases reciprocal to

{(21Ï2202), (221Ï2Ï1)} (a) and {(10ÏÏ011), (110ÏÏÏ0)} (b) Due to the higher occupancy factors

the second one (b) is used for analysis of the QPNC-14. In (a) the dark points correspond to the

projected vertices of the 4D atomic surfaces The black rhombi indicate the PAS unit cell.

B. PAS's and phononic crystals

In the next step we are going to compare the QPNC's with PNC's with their periodic

average structures (APNC's). Transmittance properties of QPNC's with low a and p„(( close

to 1 are expected to be similar to those of the corresponding APNC's in the low-frequency

part (the significance of strong Fourier coefficients for the formation of bandstructures was

demonstrated by Kaliteevski et al. [20]). The filling fractions of the PNC's with quasipenodic

and the periodic average structure are not necessarily equal, unless specifically adjusted via

changing the rod radius. A good agreement of filling fractions of the systems is equivalent

to pocc ~ 1.

(a) (b)

48

TABLE III: Vertex densities and occupancies of PAS's.

Tiling Pta PPAS1' pocc = Pt/PFAS

eightfold SQ (V2 + l)/2 (V2 + l)2/4 2/(y/2+l) = 0.83

eightfold RH (V2 + l)/2 (^+l)2/(2v^2) v/2/(v/2 + l) = 0.59

tenfold (r + l)/(sinf + r-sinf) r2/[(3 - r)2sin(27r/5)] (3 - r)/r=0.85

12-fold (2 + V3)/3 (2 + v^) 7(6^/3) 2^3/(2+ >/3) -0.93

14-fold [Fig. 5(a)] 1.25 4.94 0.25

14-fold [Fig. 5(b)] 1.25 1.30 0.96

"Vertex density of the tiling.

bVertex density of the PAS.

In this comparison a great deal of information about the way the rods are displaced or

absent from their average position is discarded, of course. There is no method that includes

the full higher-dimensional information about the infinite quasiperiodic structure, which

is projected into these deviation polygons (projected atomic surfaces), into calculations of

physical properties.

For the QPNC-8 the parameters a and pot.c suggest a high degree of averaged periodicity

and therewith a good agreement of transmittance with that of the APNC (SQ and RH). In

Fig. 6 the corresponding curves are given. The agreement is limited to the low-frequency

part of the spectra. The lower edge of the first gap as well as its central frequency are

well reproduced by both PAS's. The width of the gap (whose upper rim broadens with the

sample size) is more tightly bound by the rhombic structure than by the square one due to

the difference in occupancy. From both APNC's the QPNC differs mainly by large amounts

of vacancies, while the displacive deviations are small.

For the QPNC-10 the comparison is less clear. There is only a sharp peak left of the

second band. The edges of the first band gap can be reproduced by the APNC. Not only

is the occupancy of the PAS low, but also the displacements from it are large (aa). This

displacive and substitutional perturbation is too strong for any further transmission.

The QPNC-12 shows additionally to the low-frequency band a narrow transmission range

at the lower edge of the second band. The gap in the primitive APNC (i.e., one lattice

point per unit cell) reproduces only the lower edge of the gap. If additionally two rods are

49

introduced at positions according to the two segments of the projected atomic surface on

the longer diagonal of the PAS unit cell (see Fig. 4), both edges of the gap are reproduced.

Since these segments contribute only about half as many vertices to the tiling as do the

sections on the vertices, the diameter of the added rods was chosen half that of the others.

With these additional vertices, the displacements of the quasilattice from the PAS become

very small and because the period of the system is not changed thereby the occupancy factor

pmi remains close to 1 (however, the occupancy of the additional sites is rather low). The

QPNC-12 can be considered as an APNC with mainly vacancy disordei.

The transmission spectrum for the QPNC-14 shows even fewer cleai features than the

QPNC-10. Only the lower edge of the first gap is reproduced. In contrast to QPNC-8

and -12 this QPNC can be considered as an APNC with almost no vacancies but massive

displacive disorder. In the spectrum, there are no clear features above the first gap.

In summary, a suitable average structure of a QPNC gives information about the first

gap in its transmission spectrum (which is very important foi all applications) as well as

an estimate of transmittance in the second band. Suitable average structures have lattice

constants close to the edge length of the tiles (occupancy factor close to unity) and the

projections of the atomic surfaces are small with respect to the unit cell size. Apart from

providing information about the center and approximate width of the first gap, the average

structures can also be used to estimate its stability in case of deliberately or accidentally

introduced defects. PAS's allow one to comment on the degree of averaged periodicity of the

structure. Periodic structures are, according to Sheng [31], subject to stronger modification

of the band structure under introduction of defects than are strongly disordered structures.

This geometrical characterization of tiling is independent of the current realization of PNC's

as steel rods in water (i.e., physical systems with the structures following an eightfold tiling

are closer to periodicity on average than systems with a 14-fold tiling structure).

C. Transition structures between APNC and QPNC

The following comparison of three types of disordered PAS's of the eightfold tiling is

aimed to show how transmission changes under transition of the PAS to disordered and to

quasiperiodic structures In the first case, a statistically disordered structure {D\), the rods

are displaced from the PAS square lattice by random vectors bound by an octagon (such

50

c

Einc

500 1000

f [kHz]1500

FIG. (i: (Color online) Transmission spectra of QPNC's (black) and their corresponding APNC's

[red (gray)] agree well around the first gap. APNC bands higher than the second are absent in the

spectra of the QPNC's. Transmission through the QPNC-8 shows a stronger second band with a

lower edge, which is well reproduced by rhombic and square APNC's. The difference in the width

of their gaps is due to their different filling fractions. Transmittance through the QPNC-12 shows

a weak second band which is reproduced by the APNC only when two more rods (r — 0.25 mm)

are introduced into the PAS unit cell (see Fig. 4).

51

that the modulus of the vertex coordinates within the PAS unit cell produces the same

octagons as do the quasilattice vertices). At a displacement amplitude half of that of the

QPNC (amax = 0.135), the disordered system exactly reproduces the first band gap of the

APNC. The second band is still visible and transmittance in it is reduced from the lower

edge on. Above the second band, there is an almost constant background. The spectrum of

the system with random displacements at the same amplitude as the quasiperiodic pattern

("max — 0.27) does not show any transmitting range above the edge to the low-frequency

band, which is still the lower edge of the first band gap of the periodic structure.

In the second type, D2, perturbation of the perfect square APNC is achieved by the

introduction of vacancies. The curve in Fig. 7 results from an APNC set up by a supercell of

3x3 rods with the central rod missing. While in Dx only modifications of transmission within

the ranges of the original APNC bands evolve, the introduction of vacancies allows new

bands to form within the original gaps. These are partially closing and, for the distribution

of vacancies described, approach both band-gap edges of the QPNC. This similarity confirms

the fact that for QPNC-8 the vacancies are significant for explanations of the differences in

transmission behavior between the APNC and QPNC.

The last type, D3, combines the two previous types of disorder. Rods are positioned

halfway between each original tiling vertex and the PAS vertex closest to it. This structure

is no longer quasiperiodic but deviates from the PAS in a similar way, just to a lesser extent

(«max = 0.135). The spectrum of D3 is very similar to the one of QPNC-8 (see Fig. 7),

although the structure lacks the special long-range order. Transmittance in the second band

is reduced first in the middle of it and then at its edges. The onset of the third band of the

APNC is still visible, but beyond this range, the peaks do not seem to be correlated with

those of either the periodic or the quasiperiodic structure.

Comparison of the transmittance of the disorder types D\ and D3 shows that at a devia¬

tion amplitude amax = 0.135 the Dx system and the APNC transmit very similarly whereas

the deviation toward the QPNC (D3) at the same amplitude produces a transmittance

which is already very similar to that of the QPNC. Doubling the displacement amplitude to

«max — 0.27 does not cause much further change in Z)3 but leads to a complete loss of any

transmitting range in D]_. The range around the first band gap of D3 is almost identical

with that of the QPNC even when the displacement vectors from the APNC are reduced to

a quaiter of those in the QPNC (amax = 0.068). This agreement suggests a high robustness

52

(b) Introduction of vacancies (D2)i L. —J I i 'i I

I 1 l__ —i I i_-J

0 500,

1000 1500f [kHz]

FIG. 7: (Color online) Influence of disorder on transmittance of the APNC's [red (gray)] of QPNC-

8 is shown. In D\ (a), disorder is introduced by random displacement vectors which are bound by

the octagonal outline of the projected atomic surfaces [o/max = 0.27 (dotted)] or an octagon of half

the size [«max — 0.135 (black)]. In D-i (b), vacancies are introduced into the APNC (black). The

gap of the perfect APNC shrinks to the width of the one of the QPNC [dotted, in (b),(c)]. In Ds

(c), the deviation from the square APNC is achieved by shifting of the tiling vertices halfway to

the closest PAS vertex [amax — 0.135 (black)], creating vacancies and displacements.

of this spectral feature under structural modifications. Robustness of gaps is very important

for all types of applications of PNC's that require defects (e.g., waveguides, cavities, etc.)

without affecting the transmission properties of the surroundings.

53

II. CLUSTERS AND SCALING OPERATIONS

A. Clusters and low-frequency dips

Cluster concepts for quasiperiodic structures arc based on the observation of ubiquitously

recurring structure motifs in electron microscopic images of quasicrystals. This idea is also

supported by tiling theory. Quasiperiodic structures can likewise be described by tilings

(i.e., tiles and matching rules) and coverings (i.e., clusters and overlapping rules). So far, we

have focused on the tiling description and therewith on the length scale of an edge length

or. For the 8-fold tiling, the length scale in a cluster description is illustrated in Fig. 2. A

quasilattice is shown, which is occupied by clusters whose innermost shells consist of close-

packed eight-rings (gray circles). This quasilattice [£glust (gray lines)] is related to the original

tiling [ts (black lines)] by scaling symmetry with the factor si — (v/2 + l)2. In phononics,

clusters should be defined either by their ability to support spatially confined resonance

modes (e.g., resonances in close-packed rings [23, 24]) or by their high local scatterer density,

which makes the cluster act as a larger joint-scattering object. Such close-packed structural

motifs are prominent especially in singular tilings.

The positions of band gaps formed by scattering at such eight-ring clusters can be es¬

timated by the gaps of an APNCcl,lst. The PAS of tfust is equivalently scaled by si and

the central frequency of the first APNCclust gap should be 1/si times that of the original

APNC. At these lower frequencies the scattering cross section of a single rod is rather small.

The waves encounter only an effective medium. For a gap to open up, the cluster must

mimic a larger scattering object, which scatters more strongly at low frequencies. At the

position of the expected APNCclust gap the spectrum of the QPNC-8 has no dip but reg¬

ular Fabry-Pérot oscillations (see arrow «i in Fig. 8). Thus, the local scatterer density of

the eight-ring clusters is not sufficient to act as an inhoinogeneity disturbing the effective

medium. However, if this scatterer density is modified by a replacement of the eight-ring by

a cylinder of the same radius (QPNC D4), this inhomogeneity can be accomplished. The

first gap in the spectrum of D4 opens at the position of the first band gap of the APNCclust.

In general, a PAS of fglust is also a PAS of f8 because the two structures are scaling symme¬

try equivalent. The equivalence of an s\ times larger PAS unit cell of tg [i.e., {(Ï101), (01Ï1)}]

and the main PAS [i.e., {(110Ï), (Olli)}] of an .si-times-scaled tiling is obvious. Thus, inter-

54

. _J 1_ I \J 1 _I .

0 100 200 300 400 500

f [kHz]

FIG. 8: (Color online) Transmittancc of QPNC-8 [red (gray)]. Arrows below the curves indicate

positions of the first gap of the original APNC scaled by S3 — 2cos(tt/8), S2 = \/2 + 1, and si — s^.

In transmittance of D\ (black) the disturbance of the effective medium by the larger rods leads to

an attenuation peak at the position of the first gap of the APNC for the quasilattice scaled by .sa.

esting structures for low-frequency gaps are defined by other pairs of strong Bragg peaks of

t% (at \k\ < l/oT) or equivalently by the original PAS and the scaling and rotoscaling factors

of the tiling. In Fig. 8, potential gap positions are indicated by arrows at positions of the

first gap of the original APNC inversely scaled by further scaling and rotoscaling factors of

*8 [s1 = 2cos(7r/8) and s2 — V2 +1]. The larger the scaling factor or the larger the unit cell

of the APNCclust, the larger is the required scattering strength of the cluster for the creation

of a significant gap.

B. Scaling symmetries

Regarding the global scaling symmetry of gap positions in transmittance, a very narrow

parameter space exists for structures exhibiting this special feature. Apart from that, general

mathematical restrictions (e.g., no strict self-similarity in structures with different scaling

factors) apply, as discussed by Velasco et al [12]. In one dimension, this scaling symmetry

can be quite distinct, as can be illustrated by a Fibonacci sequence of cpoxy sheets in water

(sheet thickness 0.2 mm, separations 1 and r mm). The gap positions in the spectrum are r-

scalablc and the individual bands arc self-similar with respect to their center [sec Fig. 9(a)].

This symmetry can be understood by imagining the sequence as a tight-binding system of

resonators (space between two sheets) of widths 1 and r. The ratio of their fundamental

resonance frequencies is r/1. Thus also the positions of the bands of coupled resonators

m

2,

<uo

c

to

"Ein

0

^A

-20

\f \

^0

ts,

55

1000 2000

f [kHz]

3000

FIG. 9: (Color online) Transmission spectrum through a Fibonacci sequence of epoxy sheets in

water (a) and the same spectrum with r-scaled frequency axes [red (gray)] reveal scaling symmetry

of the gap positions. Transmittance through a diluted QPNC-8 (b) with the same spectrum scaled

by 1 + V2 [red (gray)]. A correlation between the spectra can be seen up to about 2.4 MHz.

(and gaps) can be expected to follow this rule. In 2D, the resonances of the scatterers have

nothing to do with the typical distances in their arrangement. The scaling symmetry of

gap positions has to be limited to frequency ranges in which such resonances are absent.

The degree of symmetry is illustrated in Fig. 9(b) by a QPNC-8 (r—0.3 mm, to reduce the

influence of the nature of the scatterer) whose spectrum is compared with one scaled by

1 + \/2. Despite the multiple scaling factors, the spectra are correlated up to about 2.4

MHz.

III. POLYGONAL AND STAR-SHAPED SCATTERERS

The arrangement of the scatterers clearly dominates the formation of the band structure

in the systems studied here and the properties of the scatterers will not be discussed in

detail. Nevertheless, some features of the structure can be accentuated when the scattering

objects have geometries specifically adjusted to the symmetry of their arrangement. The

cylinders have been replaced by polygonal and five-star-shaped prisms. These shapes are

56

interesting from mainly two points of view. The scattering mechanism is based essentially on

reflection of the waves at the scatterer-matrix interface. Polygonal prisms on a quasilattice

of the same rotational symmetry (or a multiple of it) form sets of broken planes, which

simplify a strong interaction of reflected waves. Apart from this, the shapes are related to

the projected atomic surfaces [see Figs. 3(c) and 4], which are an inherent property of each

quasiperiodic structure. Octagons on the eightfold quasilattice produce almost exactly the

same spectrum as the cylindrical system. If squaie prisms are used, significant reduction

of transmittance can be achieved. Especially the second band of the APNC is completely

erased. Yet the spectrum remains spiky. Pentagonal prisms forming a QPNC-10 again

yield a spectrum very similar to that of the cylindrical system. Occupying a QPNC-10 with

five-star-shaped prisms, however, leads to a considerable reduction of transmission over

broad frequency ranges. And some of the very deep gaps therein have a smooth bottom,

as is common for periodic systems (see Fig. 10). There is a clear correlation between the

spectrum in Fig. 10(b) and its r- and r2-scaled equivalents. This scaling operation applies to

both the tenfold tiling as well as the Fibonacci sequence that is formed by the broken lines

of all parallel faces of the star prisms. The correlation reaches as far as the more strongly

transmitting band around 1 MHz. This band is formed by resonance states of the single

star prism and disturbs the scaling symmetry.

CONCLUSIONS

We have performed detailed geometrical investigations of quasiperiodic tilings in order to

better understand the nature of ultrasonic wave propagation in finite phononic quasicrys-

tals with such structures. The PAS's of the quasiperiodic structures allow prediction of the

frequency ranges of band gaps of the QPNC's. The deviation of the quasiperiodic structure

from its PAS can be interpreted as a measure of aperiodicity. It is a characteristic param¬

eter for each structure, which can be used to estimate transmittance in higher bands as

well as to judge the capabilities of the different structures to form clear band gaps. This

characterization applies not only to QPNC's, but also to other physical systems with a

quasiperiodic structure (e.g., surfaces [33, 34]). Analysis of the scaling symmetries of the

structures yields information about the attenuation dips in the low-frequency band. Scal¬

ing symmetries of transmittance can be enhanced by adjusting the scatterer shapes to the

57

FIG. 10: (Color online) Comparison of the transmittance of five-star prisms on the tenfold tiling

(black) with that of a cylindrical system [red (gray)] reveals a significant drop (a). Especially above

600 kHz a deep gap with a smooth bottom opens. The spectrum is correlated (b) with r-scaled

[red (gray)] and T2-scaled equivalents (dashed) up to the strong band around 1 MHz (arrow c).

symmetry of the quasilattice. Both the deviation from averaged periodicity as well as the

tendency to form low-frequency dips are helpful for predictions of the broadband shielding

potential of a QPNC. Thus, the analysis provided here assists in the choice of a structure

for different possible applications of QPNC's (QPTC's).

Acknowledgments

We would like to thank B. Djafari-Rouhani, Y. Penncc and J. O. Vasseur for discus¬

sions and for providing the FDTD code we we adapted. We also thank Y. Psarobas and

R. Sainidou for discussions as well as Hans Reifler for technical support.

APPENDIX: HIGHER-DIMENSIONAL BASES

For the sake of completeness of the description of PAS's in Sec. I A, the bases for the

higher-dimensional embedding of the tilings are given here.

The 4D hypercubic basis dt and the reciprocal basis d* for the eightfold tiling are given

58

(in Cartesian coordinates V) by

d, — e, and2a*

with

e, = with i = 1, ,4. (A.1)

v

I cos(2ttz/8) \

sin(27T-i/8)

cos(67r?'/8)

\ sin(67rz/8) J

The physical (parallel) space is spanned by the vectors {(1,0, 0, 0), (0,1, 0, 0)}v. The length

of the 2D reciprocal basis vector a* is related to the unit tile's edge length aT by a* = l/2ar.

The 4D hyperrhombohedral bases dt and its reciprocal d* of the tenfold tiling are

/ cos(27ri/5) - 1 \

sin(27ri/5)

cos(47T?;/5) - 1

\ sin(47ri/5) /

d, = —5a*

with i = 1, (A.2)

v

and

/ cos(2ttz/5) \

sin(27ri/5)

cos(47n'/5)

\ sin(47rz/5) )

Parallel space is spanned by the vectors {(1,0,0,0), (0,1,0, ())}v The length of the 2D

reciprocal basis vector a* is related to the unit tile's edge length ar by a* — 2r2/5ar.

The 12-fold tiling can be obtained from a 6D orthonormal lattice (d basis) [28, 29], This

lattice is embedded in space such that v = {v0, w0, v1,w1,V2, w2} is an eigenbasis for cyclic

permutation of the d basis vectors. 2D parallel space is spanned by the vectors {v0,w0}.

The remaining {v„w,} span the perpendicular space. To generate the tiling and its Fourier

transform in the plane spanned by {(1, 0,0, 0, 0, 0), (0,1, 0, 0, 0, 0)}v, the lattice with basis

59

vd can be cut and projected:

v0 = (1, V3/2, -1/2, 0,1/2, - V3/2)d,

w0 - (0,1/2, -V3/2, -1, V3/2, l/2)d,

vi = (1, -A/2, -1/2, 0, -1/2, >/3/2)d)

wi - (Ü, 1/2, V3/2, -1, -n/3/2, l/2)d,

v2 = (0,l,0,l,0,l)d,

w2 = (1,0,1,0,1,0),.

(A.3)

The length of the 2D reciprocal basis vector a* is connected with the 12-fold tiling's edge

length ar by a* — l/3ar. This system is well suited for describing the atomic surfaces, but for

the derivation of the PAS it is not convenient due to ambiguous indexing of reflections. This

ambiguity is avoided when the v^1 basis is projected along its third and fourth components.

A 4D hyperrhombohedral basis d4 can then be obtained from the projected hypercubic basis

a^ by di = ai - a5, d2 = a2- a6, d3 = a3 + a5, d4 = a4 + a6.

With respect to a Cartesian basis the dt basis and its reciprocal d* are given by [32]

1

d* = ir^fäi —

em„d(i+2,4); and d* = a*eu

with

e, = for i = l, (A.4)

/ cos(2tt(z - 1)/12) \

sin(27r(i - 1)/12)

cos(10tt(z - 1)/12)

V sin(107r(/ - 1)/12) /

The 7D hypercubic basis for the 14-fold tiling is given, analogous to the 6D description

of the 12-fold tiling, by v = {v0, w0, vl5 Wi, v2, w2, d}, whereby {Vj,w,}y and additionally

d are the subspaces that are invariant under cyclic permutation of the basis vectors of a 7D

60

orthonormal lattice:

vo = (y/2/7(cos(27ri/7))7J=1)d,

w0 = (y/2/7(sm(2*i/7))7J=1)d,

vx = (y/2/7(cos(Airi/7))7J=1)d,

wi = (V2/7(svn(4iri/7))]=1)d,

v2 = {^2j7{œs^wij7))]=l)(h

w2 = (N/277(M/i(67ri/7));=1)d,

d= v^77(l, 1,1,1,1,1, l)d- (A.5)

The length of the 2D reciprocal basis vector a* is connected with the 14-fold tiling's edge

length ar by a* — 2/7ar.

IT

[9

[1«

[H

[12

[is;

[h:

E. Yablonovitch, Phys. Rev. Lett. 58, 2059 (1987).

http ://phys.lsu.edu/^jdowling/pbgbib.html

M. Sigalas, M.S. Kushwaha, E.N. Economou, M. Kafesaki, I.E. Psarobas and W. Stcurer,

Z. Kristallogr. 220, 765 (2005).

Y. Lai, Z.Q. Zhang, C.P Chan and L. Tsang, Phys. Rev. B 74, 054305 (2006).

D. Shechtman, I. Blech, D. Gratias and J.W. Cahn, Phys. Rev. Lett. 53, 1951 (1984).

R. Merlin, K. Bajema, R. Clarke, F.-Y. Juang and P.K. Bhattacharya, Phys. Rev. Lett. 55,

1768 (1985).

S. Ostlund and R. Pandit, Phys. Rev. B 29, 1394 (1984).

M. Kohmoto and B. Sutherland, Phys. Rev. B 34, 3849 (1986).

H. Tsunetsugu, T. Fujiwara, K. Ueda and T. Tokihiro, Phys. Rev. B 43, 8879 (1991).

M. Bertolotti, P. Masculli and C. Sibilia, Optics Letters 19, 777 (1994).

C. Sibilia, LS. Nefedov, M. Scalora and M. Bertolotti, J. Opt. Soc. Am. B 15, 1947 (1998).

V.R. Velasco, R. Perez-Alvarez and F. Garcia-Moliner, J. Phys.: Condens. Matter 14, 5933

(2002).

H. Aynaou, V.R. Velasco, A. Nougaoui, E.H. El Boudouti, B. Djafari-Rhouhani and D. Bria,

Surf. Sei. 538, 101 (2003).

Y.S. Chan, C.T. Chan and Z.Y. Liu, Phys. Rev. Lett. 80, 956 (1998).

61

[15] Y. Lai, X. Zhang and Z.Q. Zhang, J. Appl. Phys. 91, 6191 (2002).

[16] W. Man, M. Megens, P.J. Stcinhardt and P.M. Chaikin, Nature (London) 436, 993 (2005).

[17] M.E. Zoorob, M.D.B. Charlton, G.J. Parker, J.J. Baumberg and M.C. Netti, Nature (London)

404, 740 (2000).

[18] X. Zhang, Z.Q. Zhang and C.T. Chan, Phys. Rev. B 63, 081105(R) (2001).

[19] M. Hase, H. Miyazaki, M. Egashira, N. Shinya, K.M. Kojima and S.I. Uchida, Phys. Rev. B

66, 214205 (2002).

[20] M.A. Kaliteevski, S. Brand, R.A. Abram, T.F. Krauss, R. DeLa Rue and P. Millar, Nanotech-

nology 11, 274 (2000).

[21] D. Sutter and W. Steurer, Phys. Status Solidi C 1, 2716 (2004).

[22] S.P. Gorkhali, J. Qi and G.P. Crawford, J. Opt. Soc. Am. B 23, 149 (2006).

[23] Y. Wang, X. Hu, X. Xu, B. Cheng and D. Zhang, Phys. Rev. B 68, 165106 (2003).

[24] K. Wang, Phys. Rev. B 73, 235122 (2006).

[25] B. Djafari-Rouhani, Y. Pennée and J. O. Vasseur, (private communication).

[26] W. Steurer and T. Haibach, Acta. Crystallogr., Sect. A: Found. Crystallogr. 55, 48 (1999).

[27] A. Cervellino and W. Steurer, Acta. Crystallogr., Sect. A: Found. Crystallogr. 58, 180 (2002).

[28] J.E.S. Socolar, Phys. Rev. B 39, 10519 (1989).

[29] A. Yamamoto, Acta. Crystallogr., Sect. A: Found. Crystallogr. 52, 509 (1996).

[30] P. Repetowicz and J. Wolny, J. Phys. A 31, 6873 (1998).

[31] P. Shcng, Introduction to Wave Scattering, Localization and Mcsoscopic Phenomena, Aca¬

demic Press, San Diego (1995).

[32] W. Steurer, Z. Kristallogr. 219, 391 (2004).

[33] W. Steurer, Mat. Sei. Eng. 294, 268 (2000).

[34] Y. Weisskopf, S. Burkardt, M. Erbudak and J.-N. Longchamps, Surf. Sei., 601, 544 (2007).

62

Additional notes to article 5

The significance of strong Fourier coefficients

Wave propagation in a QPNC could be modeled by using a super-cell approach (i.e. an

approximant) and then apply the plane-wave-expansion method (see appendix B.l). In a

periodic structure the vectors of the reciprocal lattice readily designate; the Fourier coef¬

ficients important for the expansion. The same is true also for a quasiperiodic structure.

Just, these Fourier coefficients do not form a lattice. Thus, the strong Bragg-peaks can be

obtained straight forwardly from the Fourier transform and the question arises why the

simplification to the periodic average structure is necessary if the weighting of the coeffi¬

cients is known in advance? Certainly the physical quantities of a QPNC (e.g., its band

structured61 ) can be described by its stronger Fourier coefficients (which, by the way, are

often not that many). But the comparison of different structures according to a long list

of strong coefficients is not convenient. As the key parameters describing the degree of

aperiodicity, the occupancy, pocc, and the filling fraction of the PAS unit-cell, «f, seem

much more adequate. Alternatively, the ratio of the intensities of the Bragg-peaks used

to define a pseudo-Brillouin zone to the intensity of the \k\ — 0 peak could be studied.

Size dependence

The aperiodicity of quasiperiodic structures in general acts as a driving force for the lo¬

calization of waves. On the other hand, the scaling symmetry of these structures, being

a form of repetition (even though not a periodic one), pursues the formation of more

extended states. This fundamental competition has been identified early/1159 Now, ape¬

riodicity is effective in quasiperiodic systems of any size. The scaling symmetry, however,

is significant only in systems of a certain size;. Thus, the outcome of their competition is

quite crucially dependent on size.

This competition has many significant implications. Fist of all the size of the QPNC

becomes important for the interaction of waves. Finite QPNC's are likely to act like

disordered PNC's because the scaling-symmetry may require huge sections of a tiling to

be used. This may be a problem, because the chosen sections does not represent the whole

quasiperiodic structure. On the other hand, any application of QPNC's is doomed to use

such finite sections and therewith inherit only part of their properties.

The mean free path lengths, /inft„ can also be affected by this contest. If the disorder¬

like aperiodicity is favored by the chosen section, then /mfp is supposed to decrease and

if the scaling symmetry is pronounced it is supposed to increase. With decreasing Zmfp

wave propagation is expected to become more diffusion-like. And therewith, the spectral

features (e.g., bandgaps or defect states) lose their sensitivity to structural modifications

in their not to near surrounding because of their strongly limited range of coherency (see

63

co

400 600

f [kHz]

1000

Figure 2.8: The evolution of transmittance with the sample size. In red (grey) line

transmittance through a 9thgeneration Fibonacci sequence and in black through a 3rd

generation sequence is shown.

Shengdm). Localization, delocalization and their competition is characteristic for every

quasiperiodic PNC's.

In PNC's, typically eight to twelve unit-cells in direction of transmission are sufficient

to draw a clear picture of the band structure of the crystal. In the section of the 8-fold

QPNC in articles 4 and 5 (Sec. 2.2 and 2.3) the number of close-packed eight-rings is

comparable (and the structure is larger than 10 PAS periods). We therefore assume, the

chosen sections are representative for the quasiperiodic structure. To further elaborate

on this point, the size dependence of a ID Fibonacci QPNC shall be shown.

In Fig. 2.8 transmission spectra of ID Fibonacci QPNC's consisting of epoxy sheets

in water are shown. The red (grey) curve represents a 9lh order and the black curve a

3nd order sequence. Quite evidently, the shorter sequence already produces all the major

gaps in transmittance. But, of course, the fine structure (with the self-similarity) is not

observed for such a short sequence. Ultimately, there is no convergence. For the fine

structure a significant level of detail has to be specified. But the strong band gaps remain

the same.

Implications for real quasicrystals

To stress the concept of PAS and band structures to the outmost a note on the stability

of real quasicrystals shall be given. Given, quasicrystals are stabilized electronically via

Hume-Rothery mechanism (i.e. the structure adjusts in such a way, that the electronic

band structure, which directly depends on the structure, of course,dl4(i allows a unique

distribution of electrons with a global minimum of the total energy). A structure, which in

general supports the formation of deep and wide gaps is more likely to allow a minimization

of this total electronic energy. While there are experimental findings of quasicrystals with

8-, 10- and 12-fold symmetry, for the 14-fold symmetry nothing has been found. This

could be due to the very aperiodic nature of the underlying structure.

64

2.4 Icosahedral phononic quasicrystals

D. Sutter,* P. Itten, P. Neves and W. Steurer

Laboratory of Crystallography, Department of Materials, ETH Zurich

(February 2, 2007)

Abstract

Periodic average structures (PAS) were found to assist the prediction of the band struc¬

tures of quasiperiodic phononic crystals (QPNC's) in two dimensions. In this first report

on a three-dimensional QPNC we present experimental transmission spectra of an icosa¬

hedral QPNC as well as a phononic crystal with its PAS. The transmission spectrum of

the QPNC features a very distinct band gap. This band gap agrees perfectly with the one

of the PNC in position and width. The suitability of the periodic average structure to

characterize; quasiperiodic structures and to estimate their physical properties is further

confirmed herewith.

'Electronic address: daniel. sutterOmat. ethz. ch

65

Research on quasiperiodic heterostructures has taken thirteen years (1985<i8' -1998d19)

to make the step from ID to 2D. And it took another eight years for the third dimension

(Man et al.AM). This leisurelincss in an otherwise fast developing field has many reasons

but mainly reflects the advances of production techniques for such structures on the

nanoscale (in photonics). Generally, investigations of 3D QPNCdG4, dM did not reveal

any physical properties, fundamentally different from those of 2D QPNC. Band gaps were

found in all systems but hard scattering regimes proved again to be more difficult for

obtaining clear bands and gaps.d-64 Nevertheless, then; are some interesting questions

regarding the distinctiveness of gaps or the tendency for wave-localization in 3D QPNC's.

Beyond this, of course, the fabrication of such systems is a challenging. Possible techniques

are the microwave stereolithography,d-8f) holographic methodsdii2 or inversion methods/177

This present study reports on the first 3D QPNC. The interpretation of the transmission

properties of the icosahedral QPNC (i-QPNC) is aided by using the periodic average

structure. This approach has been successfully applied to various 2D QPNC's before (see

Articles 4 and 5 in Sec. 2.2 and 2.3).

The icosahedral structure and its PAS

The icosahedral tiling,d78, d131or three-dimensional Penrose tiling has been frequently

used as a model quasilattice of real quasicrystals. While many different tilings have been

thoroughly analyzed in 2D, in 3D the icosahedral one is the very representative. For the

calculation of the vertices of the tiling we used the code by Janot,d-5/| which bases on the

higher-dimensional description (see appendix C). The 6D basis of the tiling (taken from

Steurer and Haibachd137) is given by

di1

2a*

/o\

0

1

0

0

V1/

and d, = —2a*

( sin0cos(27r?'/5) \sinÖsin(27rz/5)

cos/9

—sin0cos(47ri/5)

—sin0siii(47ri/5)—COS#

for i = 2, (2.G)

with 9 = 63.44° and a* = l/2ar (ar is the edge length of the unit building blocks).

The atomic surfaces are regular triacontahedra. Projection of the 6D lattice and the

triacontrahedra along the three directions (ïllulO)^, (OlIlOÏ)^, (Ï001Ï1)D onto parallel

G6

Figure 2.9: The infinite isocahedial tiling modulo the unit cell of the PAS results in

point&ets circumscribed by tiiacontahedra on a fee lattice (a). In (b) a photograph of the

z-QPNC (consisting of steel beds in polyesther) is shown during fabrication.

space leads to a PAS spanned by the cubic basis

1 \ / 0 \1

_...1

tan(7r/5)„pas_

„PAS_

An —

2a~*\ 0

and a?AS =

/

2a*

which has a lattice parameter a£AS = 2tan(7r/5) • ar. The PAS lattice is occupied in a

face-centered (fee) fashion by undistorted triacontahedra. They fill af = 19.5%) of the

volume of the unit cell [see Fig. 2.9(a)]. The maximal deviation of a tiling vertex from the

PAS is amax = 0.26 • aPAS. The occupancy of the projected atomic surfaces is pocc — 1.18.

Thus, when the PAS and the tiling are supeiposed, 18% of projected atomic surfaces

contain two veitices of the tiling. The icosahedral tiling can be constructed with two

rhombohedra, whose faces are equivalent rhombi. The closest distance between vertices

in the tiling is the shortest diagonal of the prolate rhombohedron dmm — 0.563 • a,

Construction of the z-QPNC and measurement of ultrasonic trans¬

mission

The 7-QPNC was fabricated in a somewhat inconvenient, but very accurate manner. Into

a block of polyester, holes were drilled at the vertices of a layer of the tiling with two

specific and proximate1 z-coordinates. To avoid air inclusions, these holes were first wetted

with liquid polyester and subsequently equipped with steel balls of 1 mm diametei. The

67

0

m

f 20

SE(A

200 400 600 800 1000 1200

f [kHz]

Figure 2.10: Measured transmission spectra of the ?-QPNC (black) and its APNC [red

(gray)] arc shown. The first band gaps coincide perfectly.

whole array of positioned spheres was then doused with a thick layer of resin. After

curing the resin (24h at 40°C), the thick layer was machined down to slightly above the

z-coordinate of the following set of vertices, at which again spheres were positioned in

the same1 way. The phononic eiystal with the PAS (APNC) was fabricated equivalently.

Due to the complete curing of each individual layer, a weak interface remained between

them. This interface does not feature an acoustic impedance mismatch and is therefore

not relevant.

The critical step in the fabrication was the drilling of nearest-neighbor holes, which

required a minimal separation of the vertices in projection to the1 xy-planc. Therefore,

the vertices of the icosahedral tiling were scaled to dmm — 1 'zo mm, with dmm the closest

distance between two tiling vertices. This adjusts ar to 2.22 mm and «pas to 3.227 mm.

The filling fraction of the APNC is then 0.062 and the one of the QPNC is 0.074. A

picture of the ?-QPNC under construction is shown in Fig. 2.9(b).

The transmission spectra were measured with two Panametrics transducers with cen¬

tral frequency 800 kHz (driven by a Panametrics 5072PR Puiser). The signals were dig¬

itized using a LeCioy LT354 digital oscilloscope and recorded with a PC. The resulting

plots are shown in Fig. 2.10(a). The APNC is 9 x 9 unit cells in the plane perpendiculai

to and 4.5 unit cells in direction of transmittance (1458 spheres). The QPNC consists of

3438 spheres. In ordei to check the whole measurement procedure and also to get the

full band structure of the APNC, multiple-scattering simulations were performed using

the ADUT package/1 m The material parameters used for the polymer are q = 2000 m/s

ct = 1230 m/s and p = 1220 kg/m2.

Results and discussion

First of all, the i-QPNC shows a deep and large band gap (see black curve1 in Fig. 2.10).

This gap is much moie distinct than those of the 2D QPNC's. There are several aspects

1 _j I i 1 ] L

68

ü

o

(D

3

0

.160 -80 0 0 0.5 10 0.1 0.2

(a) (b) (c)Transmission [dB] Re[k]a0 In lm[k]a0 /jt

Figure 2.11: Transmission spectra (a) of the; /cc-APNC in experiment [red (gray)l and

simulation (black). In the; real part of the band structure (b) the first transmission gap is

occupied by a deaf band. There is only a very narrow complex band visible at the lower

edge of the gap (c), which draws nicely the first ripple of transmittance at the lower gap

edge.

relevant to this fact. First, the acoustic impedance contrast between the polymer matrix

and the steel spheres is smaller than the one in the steel/water QPNC's. Thus, the

scattering strength of the spheres is smaller. Second, the filling fraction of the sphere-

crystal is rather low, i.e., the size of the spheres is small relative to the lattice constant.

Thus, at oxZr/Cmatrix ~ 1, where strong Bragg scattering (i.e., first gaps) is expected,

the scattering strength of the spheres is low. Tin; spheres act more or less like point-

scatterers. Compared to the 2D QPNC's of the previous section these 3D (Q)PNC are

not too strongly scattering systems. And last but not least, in 2D, the localization of

waves can be achieved by any defect no matter how weak its perturbing potential is.

In 3D, localization can be achieved only if these potentials exceed a certain threshold

level (see appendix A.3 or Economou'131). All these points suggest a smooth variation

of transmittance, with not too broad but sharply bound gaps. On the other hand, in

the presence of the two additional transverse polarizations, the width of the gap appears

remarkable.

The first gaps of QPNC and APNC agree very well in position and in width. This

is backed by the small deviation of the tiling from its PAS. The displacements of the

vertices from the PAS, amax and a/, are small, similar to those in the 2D 8-fold and

12-fold tilings. The occupancy of the PAS vertices is larger than one, i.e., there are no

vacancies to be considered. Accordingly, the edges of the first gaps of QPNC and APNC

are nearly identical.

The simulated transmission through the APNC is shown in Fig. 2.11 (a). The positions

69

-

.__.- -JW

- ^_j~" -

—s--*

~~^~>e^___

-*

-

^

CD

TJ.

m

c

g

-160 -

I ' fII 1Il II

! il

^vx

1

1

1

1

1

jï)/ilII

! '

I !1 '

1 III '

II II

[mil ! /!\ i / i

[100],

l

1

1

1 [110]! !

\

I ' I 1

(c)

«1rV'

liF 11!"

H"Vi t

^/Jv

2-fold 5-fold

Ai

'

it* %

i

111

i 1 '3-fold

' '"||

300 400 500

f [kHz]

o

(d)

400 800

f [kHz]

-40

Figure 2.12: Deviations of the measured and simulated spectra can be explained by

absorption on the one hand and air inclusions around the steel spheres on the other.

Absorption (a) in the polymer [red (gray)] leads to a smooth reduction of the maximal

transmittance with the frequency. If there are air spheres [red (gray) solid] instead of

steel spheres in the PNC [red (gray) dotted] the first gap vanishes but the second one

remains (b). Thus, air inclusions around the steel spheres explain the depth of the second

band in experiment. The overlap of the first band gaps of the fee crystal is very small

[shaded area in (c)]. the lower edge of the gap in [100] direction exactly coincides with the

upper edge of the gaps in [il0] and [111] directions. Measurements of the transmission in

direction of the 2-fold, 3-fold, and 5-fold axes of the i-QPNC revealed a large overlap of

the first band gap, as expected.

of the gaps agree well with the experimental findings. The band structure reveals that

within the first transmission gap there actually is a band [Fig. 2.11 (b)]. This so called

deaf band, however, does not contribute to transmission because its states cannot be

excited by the incident wave field for symmetry reasons. The deaf band covers the upper

part of the gap. At the lower edge a complex band of evanescent states (see appendix

A.3) is shown, which fits perfectly to the logarithmic transmittance [see Fig. 2.11 (c)[.

In Fig. 2.12(a) the influence of absorption in the polymer matrix on transmittance is

shown. If an imaginary part, 195 i m/s, is added to q and ct in the simulation the gentle

reduction of maximal transmittance with the frequency, as observed in experiment, can be

seen. Furthermore, while the first gap is much deeper than the second in the simulation,

the experimental curve displays just the; opposite. The measured depth of the first gap is

limited by the limited resolution of the receiving transducer in this range. The depth of

the second gap seems to be due to the overlap of gaps of the APNC with steel spheres and

air spheres respectively [see Fig. 2.12(b)], Thus, the air inclusions which can be observed70

around a small fraction of spheres should support formation of a deep second gap and the

first one is underestimated due to lack of better resolution. Absorption, air inclusions as

well as limitations of the experimental setup can be expected to affect transniittance of

the i-QPNC in the same way. The isotropy of transmission of the two PNC's differs very

decisively. While the first band gap of the APNC in directions of the 2-fold and 3-fold

axes does not overlap with that along the 4-fold axes at all a narrow omnidirectional gap

opens at the upper edge of the latter [Fig. 2.12(c)]. The measurements of transmission in

different directions of the z-QPNC is rather difficult due to the oblique angle of incidence

on the sample in all but one direction. Nevertheless, along the 3-fold and 5-fold axes the

first gaps are even larger and coincide very well with the gap measured in direction of the

2-fold axes [Fig. 2.12(d)].

Conclusions

The investigation of an icosahedral phononic quasicrystal has shown that the periodic av¬

erage structure is very helpful for the prediction of physical properties of 3D quasiperiodic

systems. 3D QPNC's (QPTC's) have hardly been treated so far. The clear focus on 2D

structures can be justified by the potential applications especially of photonic quasicrys-

tals in optical devices, which ultimately are supposed to constitute the optical computer.

In optical devices the typical architectures are still planar. With increasing degree of

integration, however, the need for 3D QPTC's could arise1. More generally, 3D QPNC's

and QPTC's can be expected to act more similar to their periodic counterparts because

of the reduced tendency of wave-localization as well as the lower filling fractions. This

could ease working with such structures.

71

2.5 QPNC's and clusters (Article 2)

The following reprinted article deals in more detail with QPNC-10's based on Penrose

tilings with different decorations. The occurrence of close-packed ten rings is very different

in these; patterns. The influence of these rings on the formation of band gaps as well as

on the localization behavior of the waves is discussed.

D. Sutter and W. Steurer, Phys. Status Solidi C 1, 2716 (2004).

72

phys. stat. sol. (c) 1, No. 11, 2716-2719 (2004) / DOI 10.1002/pssc.2004Q5398

Ultrasonic investigation of phononic Penrose crystals

Daniel Sutter* and Walter Steurer

Laboratory of Crystallography, ETH Zurich, Wolfgang-Pauli-Strasse 10, 8093 Zurich, Switzerland

Received 15 July 2004, revised 9 August 2004, accepted 27 September 2004

Published online 17 November 2004

PACS 43.40.Cw, 62.30.+d, 71.23.Ft

This article reports on experimental ultrasound studies of two-dimensional phononic quasicrystals consist¬

ing of steel rods submerged in water. The arrangements follow Penrose tilings with different decorations.

Transmission spectra reveal no gaps indicating band gaps in the system, however, characteristic dips were

found for the different arrangements. Three tilings were considered. Two of them have the same unit-tiles

(same short range order), two of them the same arrangement (same long range order). In order to investi¬

gate the contributions of local arrangements to the spectral properties of the whole structure, several clus¬

ter motifs and rectangular subsections of the phononic quasicrystals were studied.

© 2004 WILEY-VCH Verlag GmbH & Co. KüaA, Weinheim

1 Introduction

In the last ten years a new interest in elastic wave propagation and localization has been prompted by

investigations of phononic crystals. Conventional (periodic) structures of this type have been investi¬

gated extensively, and the first steps towards quasiperiodic structures in one and two dimensions have

been reported (e.g. Velasco e* al. [1], Lai et al. [2] and Suiter et al. |3J). Apart from these, several papers

on photonic quasicrystals have been published [4, 5]. Zoorob et al. [6] and Bayindir et al. [7] have found

manifestations of bandgaps in transmittance measurements of their photonic structures experimentally.

Most of the published data concerns bandstructures with large gaps, which are of widths and depths

comparable to those of periodic structures. Others report on structures featuring no gaps but only de¬

pleted regions in the density of states [5]. The question as to whether or not there are gaps in bandstruc¬

tures of photonic quasicrystals is therefore positively answered. What remains is to categorize the struc¬

tures according to the existence of bandgaps and to investigate the mechanisms, which create them. Sev¬

eral proposals that address the question of possible mechanisms have already been made in literature on

photonic quasicrystals. The seminal work of Chan et al. [4] contained some of them. They investigated

the dependence of the transmittance on the extension of the crystal sample and found that the spectral

gaps can be found in rather small subsections of a structure. Their conclusion was that local scattering is

governing the formation of bandgaps rather than global scattering and that therefore long-range periodic

order is not a prerequisite for the existence of a gap (see also Hase et al. [8]). Wang et al. [9] then dis¬

covered, that there are sharp peaks in transmittance spectra of defect-free photonic quasicrystals, which

are associated with modes strictly localized at singularities in the structure. It was suggested that these

modes propagate via hopping from one singularity to another. The density and distribution of such po¬

tential sites for wave localization were found to have a strong effect on transmittance properties.

The first published work on two-dimensional phononic quasicrystals (Lai et al. [2]) did not go this far,

and only reported on the existence of gaps in the bandstructure of the investigated quasiperiodic square

triangle tiling. To our knowledge, no-one outside our group [3], has yet performed according experi-

"

Corresponding author: e-mail: [email protected] ,Phone: +41 1 632 37 27, Fax: +41 1 632 11 33

» 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Wcinhdm

73

phys. stat. sol. (c) 1, No. 11 (2004) / www.pss-c.com2717

ments on phononic quasicrystals. In this article we report on the experimental characterization of trans¬

mittance properties of phononic Penrose crystals and their clusters.

2 Experimental details

Phononic quasicrystals were realized as arrangements of steel rods (1 mm diameter) submerged in water.

The liquid matrix medium allows the evolution of transmittance with the sample shape to be studied

efficiently, by occupying arbitrary sections of the quasilatlice (defined by a bore pattern in parallel Mo

sheets) with rods. For the characterisation of the crystals, ultrasonic transmission spectroscopy was used.

The ultrasound pulses were generated by a Panametrics 5072PR pulser-receiver and captured by a Le-

Croy LT354 digital oscilloscope. The spectra were measured al 20 slightly different positions and aver¬

aged to obtain information more representative of the whole non-periodic structures (see also Ref. [5]).

This causes some weak and narrow spikes to disappear but enhances the stronger peaks in the spectra.

The Penrose tilings, on which the three crystals are based, are shown in Fig. 1. The first one, (Fig. 1,

PCI), is a singular rhombic tiling, producing an inhomogeneous crystal with close packed ring structures

(which, according to Wang et al. [9], could be potential sites for localization). The second crystal (Fig. 1,

PC2) is based on a kite-and-dart tiling. The number of densely packed ten-rings of rods is much higher in

this structure, but the overall homogeneity is better. The third tiling (Fig. 1, PC3) is again a rhombic

tiling and can be obtained from PC2 by substitution of the unit tiles. The tilings were generated using the

generalized-dual-method [10]. The three tilings were chosen to investigate the influence of decoration

and of the actual arrangement on the spectral properties. Tilings one and three are both rhombic tilings

and have similar short-range order. The tilings, however were generated from different dual grids. The

differences are obvious. Tilings two and three are closely linked by the tile substitution rules. The unit

tiles (short range order) are different but the global arrangement is similar. The fraction of points that are

common to all point sets after scaling the tilings with edge-lengths of the tiles are 0.39, 0.62 and 0.40 for

the three patterns respectively. The filling fractions of the structures are 0.35,0.42 and 0.28, respectively.

3 Results and discussion

Transmittance curves for all three structures, and various substructures, \,ere collected (0.4-3.5 MHz). In

figure 1, transmittance through the three different crystals is compared for the direction of wave propaga¬

tion perpendicular to one of the five star axes of the dual grid. The spectra consist of a large number of

narrow peaks and spikes. Neither broader ranges of strong transmission nor continuous ranges of strong

attenuation, indicating bandgaps of the systems, can be seen. Transmittance in general, is very low for

waves with frequencies above the edge to the long-wavelength-limil between 0.3 and 0.4 MHz. Three

broader minima are indicated with arrows in each spectrum. These three dips appear at approximately the

same relative frequencies, if the spectra are compared on a frequency scale normalized with the velocity

of sound in water and the minimal bore-to-bore distance of the patterns (1.05, 1.20 and 1.15 mm, respec¬

tively). For the structures PCI and PC3 this distance corresponds to the short diagonal of the skinny

rhomb. In the kite and dart tiling this distance corresponds to the short edge of the tiles, which is, by

definition of the substitution rules equal to the edge of the rhombus of the other tilings, which is 1.618

times (i.e. t times) longer than the short diagonal of the skinny rhomb. Thus the whole tiling PC2 is

scaled differently from the others, but the local arrangements are scaled analogously (e.g. the close

packed rings of ten rods). Since these three broad dips agree in their position on this relative frequency

scale, they must be due to local arrangements, rather than to the global structure. This is also supported

by the series shown in Fig. 2, which illustrates the evolution of transmittance with the thickness of a

rectangular section of crystal PC3. Even for a thin slab of the structure, sharp attenuation peaks appear at

certain frequencies. In periodic crystals they would grow in depth roughly exponentially with the system

size (e.g. an analogous hexagonal steel rod crystal in water with filling fraction 0.4, features a gap from

0.6-0.95 MHz along l~J [3]). Here, nothing similar can be observed. Transmittance in general is being

reduced, but the relative depths of the dips in the curves are retained. The remaining ranges of stronger

transmission become fragmented. The resulting spikes do not converge towards a stable shape, but can

© 2004 WII.EY-VCH Verlag Cm bH & Co. KGaA, Weinheira

74

2718 D. Sutter and W. Steurer: Phononic Penrose crystals

change with modest modification of the sample shape. The diagram clearly demonstrates the absence of

extended bandgaps. The three dips mentioned above emerge as characteristic features of all curves.

PC3

*"^**W»mUPC2

PC1

soo 1000 15O0

f [kHz]

2O00 2500 3000

Fig. 1 Comparison of transmittance through the different phononic Penrose crystals. Three broad minima are high¬

lighted with arrows in each curve. On the right hand side the original tilings and patterns marking the positions of the

rods in the crystals arc shown. Direction of transmission for the spectra is vertically through the given sections.

.•."•vi-'.v;;

ra 10-

Vi-.V

3500

Fig. 2 Transmittance through rectangular sections of the phononic Penrose crystals PC3. Direction of wave propa¬

gation is vertically through the sections shown on the right hand side. The formation of the three characteristic dips

in the spectrum of phononic Penrose crystals is clear.

If the positions of these dips on the relative frequency scale are determined by local arrangements,

then it would be interesting to know which motif is responsible for their creation. In Fig. 3a, transmit¬

tance curves for various small clusters of rods contained in PCI and PC3 are given. The formation of the

three dips as a function of the size of the cluster can be clearly seen (i.e. C21 (consisting of 21 rods) and

CAT). In other clusters, the original dips do not get deeper with the size of the structure but vanish upon

increasing size and only appear again for large structures. In Fig. 3b transmittance through a Gummelt

cluster as a function of direction of transmission is shown. At a 90° angle the lower two dips disappear,

and a new dip between the other two is visible. The dip at 2.25 MHz remains (in the cluster spectra these

dips appear at lower frequencies than in the one of the large structure). This can also be observed in the

spectra of the full crystals. The lower dips are anisotropic, while the one at 2.25 MHz remains open for

all directions of wave propagation. Also several motifs of PC2 produce dips. It is therefore not possible

to associate the formation of these gaps with a specific structural motif. Since a large number of motifs

O 2004 WILLY-VtH Verlag dm hH & Co KGaA, Weinheim

75

phys. stat. sol. (c) 1, No. 11 (2004) / www.pss-c.com2719

seem to produce these spectral features, the question arises as to whether the very fundamental unit tiles

(i.e. the rhombus, kites and darts) are sufficient to determine the position of the dips on the frequency

scale. This question could be pursued further by studying the periodic approximanl structures of these

unit-tiles experimentally.

i . i i i I L, il- i J i . II « Ulli ill

500 1000 1500 2000 2500 3000 500 10O0 1500 2000 2500 3000

f [kHz] f [kHz]

Fig. 3 In diagram a, the formation of the three dips for PC 1 and its clusters is clearly present. In diagram b, trans-

mittance through PC3 and its clusters are shown. The formation of the lower two of the dips can be observed (grey

bars). There are also indications of the dip at the highest frequency of the three, but they are shifted to lower fre¬

quencies. The topmost curve (taken at 18°) also exhibits a peak there (isotropic dip), whereas the lower ones arc

replaced by a single minimum between the others (anisotropic dip).

The crystals PCI and PC2 contain very close packed clusters (e.g. featuring CI 1), which appear to the

eye as inhomogeneities (see Fig. 2). Such motifs have been identified as those responsible for localized

modes in photonic structures (see Wang et al. [9]). PC3, in contrast, consists of a rather homogeneous

arrangement of rods. The fact, that the lower dips are obscured by sharp peaks for large samples of PC2

could be connected with the presence of coupled, localized modes propagating from one such cluster to

the other. Further clarification of this suggestion should emerge from theoretical calculations.

4 Conclusions and outlook

We have shown the transmission properties of different phononic Penrose crystals. No clear evidence of

bandgaps was found but all of them exhibit broad dips in the spectra at the same relative frequencies

(normalized with minimal bore-to-bore distance of the patterns). Some dips are anisotropic and one is

isotropic. The same dips were found in transmittance curves of several clusters of the crystals featuring

the same anisotropic behaviour. The structures of the different clusters creating the characteristic dips are

quite different. A further localization of the origin of these dips will be attempted by calculating the

displacement field maps und by experimental studies of periodic approximants.

References

[1] V. R. Velasco and J. E. Zàrate, Prog. Surf. Sei. 67, 383 (2001).

[2] Y. Lai, X. Zhang, and Y. Q. Zhang, J. Appl. Phys. 91, 6191 (2002).

[3] D. Sutter, G. Krauss, and W. Steurer, Mater. Res. Soc. Symp. Proc. 805, 99 (2004).

[4] Y. S. Chan, C. T. Chan, and Z. Y. Liu, Phys. Rev. Lett. 80, 956 (1998).

[5] M. A. Kalitevski, S. Brand, R. A. Abram, T. F. Krauss, R. De La Rue, and P. Millar, Nanotechnology 11, 274

(2000).

[6] M. E. Zoorob, M. D. B. Charlton, G. J. Parker, J. J. Baumberg, and M. C. Netti, Nature 404, 740 (2000).

[7] M. Bayindir, E. Cubukcu, I. Bulu, and E. Ozbay, Europhys. Lett. 56, 41 (2001).

[8] M. Hase, H. Miyasaki, M Egashira, N. Shinya, K. M. Kqjima, and S. Uchida, Phys. Rev. B 66, 214205 (2002).

[9] Y. Wang, X. Hu, X. Xu, B. Cheng, and D. Zhang, Phys. Rev. B 68, 165106 (2003).

[10] G. G. Naumis and J. L. Aragon, Z. Kristallogr. 218, 397 (2003).

© 2004 WILEY-VCH Verlag Gm bH & Co. KGaA, Wcmheim

76

Additional notes to article 2

Tight-binding cluster resonances

We have performed experiments with multiple clusters in the QPNC-8 (see article 5 in

Sec. 2.3), the kite & dart QPNC-5 and also with QPNC-12, which features very closely

packed rings (see Fig. 2.13). In none of these systems we have found any sign of res¬

onances of such rings consisting of hard steel rods. In photonics such ring resonances

have been found in several systems. Rockstuhl et al.lLln show that the cylinders in their

QPTC's clearly have strong resonances. The same was found by K. Wang,d-157 who also

explained their hybridization to cluster resonances. Wang et fli.fL159' d,16° do not refer to

cylinder resonances as the origin of cluster-resonances. However, their QPTC's feature

lower contrast of dielectric constants and higher filling fractions than those of the other

two groups mentioned (i.e. large cylinders relative to ar and low wave velocities in them

both shift the rod resonances to low frequencies). It is therefore reasonable to assume

that these localized modes are actually cluster resonances too. The question whether or

not there can be such ring resonances under absence of the single-rod resonances remains

open.

.40 I u__- 1 . —I i 1 J 1- ' '

400 600 800 1000 1200

f [kHz]

Figure 2.13: Comparison of transmittance measurementsthrough a quarter slab of QPNC-

12, one containing the origin, the other one not. Neither at the pseudogap around 600

kHz nor at the first PAS gap around 400 kHz there is any sign of a resonance peak in the

spectrum of the slab with the 12-ring.

Amplitude-map analysis

The visual inspection of the wave functions is not a fruitful approach, however tempting

its simplicity may be. The sheer amount of information of an amplitude map of the

total wave-function may not feature any pronouncement of localized states or even give

a clear picture of the general attenuation (mean free path). The ballistically propagating

intensity may be covering weaker effects (and its reduction is not trivial).

77

Chapter 3

Synthesis

From the discussions in the preceding articles we want to draw four major conclusions

and propose issues for further investigations.

Major conclusions (recommendations)

1. in resonance-based systems, quasiperiodic structures are very favorable. Their ro¬

tational symmetry clearly induces isotropic transmission properties. Bands and

gaps form analogous to those in PNC's just more isotropic. We have identified no

draw-back for using a QPNC-14 for the fabrication of an acoustic or optical de¬

vice. However, already the 8-fold tiling produces an almost completely isotropic

first bandgap. Higher rotational symmetries are only necessary to achieve the same

for the higher gaps.

2. in strongly scattering systems without resonances, on the other hand, quasiperiodic

structurels seem less favorable. Among all quasiperiodic structures analyzed, the

most periodic one (on average), QPNC-8, produces the clearest spectral features.

Given the constraint of a high-contrast materials combination for the construction

of a QPNC- or QPTC-based device, a maximization of isotropic transmission prop¬

erties can be achieved only at the expense of clear bands and gaps.

3. especially in such cases the periodic average structures (PAS) of the quasiperiodic

tiling is very helpful tool for the prediction of physical properties of quasiperiodic

systems. This approach is applicable to all quasiperiodic structures and since it

is based on Bragg scattering its validity restricted to QPNC's (QPTC's) without

strong resonances of the single scattering.

4. SD Bragg-scattering QPNC's seem to be very convenient to work with. Because

they generally have lower filling fractions and because 3D waves are intrinsically

less prone to undergo localization, 3D QPNC's can be expected to act more similar

to PNC's and produce less chaotic transmission spectra than 2D QPNC's do.

78

Proposals

1. Experimentally, a very helpful thing to do is to measure the mean free path of

wave propagation. As discussed in Sec. A.3 this requires spatially resolved intensity

measurements.d8ü Such experiments could shed more light to the propagation of

critically localized waves in QPNC's (see Delia Villa et al.d-25).

2. Computationally, a revision of the FDTD-code implementing advanced meshing

techniques would provide the basis for the investigation of far larger samples. In¬

vestigations of large samples would favor the influence of the scaling symmetry of

quasiperiodic structures in its contest against aperiodicity.

3. With respect to applications, searching for strong resonators usable for QPNC's is

of prime importance in the light of the above recommendations.

79

Appendix A

Theory of elastic waves

In this first appendix the general theory of clastic waves and their propagation are sum¬

marized. The fundamental elastic wave equation will be discussed as well as the scattering

from solid rods in liquid hosts (Sec. A.2). Finally typical tiansmission and attenuation

legimes (types of wave functions) will be analyzed (Sec. A.3). The subsequent chapter

will then show, how this is all treated by the different computational methods (Sec. B).

A.l The wave-equation

In this section we derive the elastic wave-equation for non-homogeneous systems.tiM'a °

By non-homogeneous we understand a composite of volumes of homogeneous elastic bod¬

ies. The volumes of these1 homogeneous regions arc large in phononic crystals and therefore

there is no need to consider any aspect of atomicity. The wave-equation links the spa¬

tial distribution and the temporal evolution of the elastic stress and the elastic strain in

the materials and along their interfaces. The interaction of stress and strain govern the

propagation of elastic waves. The wave functions solving the wave-equation contain all

information about the interesting propeities of a phononic crystal (e.g., transmittance,

dispersion, local amplitude distributions etc.).

The wave-equation is a simple combination of Newton's law of motion and Hooke's

law. The tensor of the elastic tension oi stress er (present at a certain point in the system

r at a ceitain time t) is given byAF

.= - (A,)

a directed foice AF acting on an area element a with normal vector n. This area clement

then suffers the tension

/ orjnx + axyny + a^znz

(A.2)

80

The stresses on all faces of a volume element exert a total force F on it

F= [ ada. (A.3)Ja

Using the Gaussian theorem F transfers to

F% = J (alxnx + aiyny + alznz)da = J (-^ + -^ + -£f)dV. (A.4)

According to Newton's law of motion the net force F, resulting from the stresses at all

faces, accelerates the movement ù of a volume element dV about its initial position

F, = / pihdV (A.5)Jv

and therewith

/^ +^ + ^y (A6)V ox ay oz J

The stress tensor a is related to the strain tensor £ via Hooke's law

(Tij — 2_^ CijklEkl (A-7)

and e can be defined by the spatial derivatives of u (assuming small, elastic displacements,

i.e. neglecting second order terms)

£ki —1 /duk

,

duil^ + F-). (A-8)\OXl

ax* /2 VUX{ U£h

The tensor C relating stress and strain is called the elastic stiffness tensor and is of

fourth rank. In crystalline materials it induces the symmetry of the structure to the

physical property stress, it is then anisotropic and can have maximally 21 independent

parameters (triclinic crystal systems). For isotropic, homogeneous media it contains only

two independent parameters (e.g., A and p) and Eqn. A.7 becomes

a,j = 2fi£t] + XÔV ^2 Ekk- (A.9)k

The Lamé constants A and ß are connected with the longitudinal and transverse speeds

of sound by

ci = \/(\ + 2/y)/p

ct = JvTp (A-10)

Stepwise substitution of Eqn. A.9 into Eqn. A.6 gives

p =

d(7xx+

da*v+

d(Txz(A.11)

dx Oy dz

81

with

<Trx — (2p + X)Exx + H£!JV + £zz)

^xy £p&xy

Gxz = 2p£xz (A-12)

d2ux f d2uy d2vz

dx2 \dydx dzdx.

/d2ux d2uy\ /d2ux d2uz \

V dy2 dxdyJ V dz2 OxdzJ

/d2Ux d2ux d2ux\ x(d2ux d'2uy dV \

V dx2 dy2 dz2 / V dx2 dydx dzdx)

Mtt + t^ + wt-)- (a-13»\ ox1 dxdy dxdzJ

Combining the other two likewise derived components of the force F leads to the vector-

notation of the wave-equation for elastic waves in homogeneous media

F = fiV2u + (A + /z)V(V • u) = p~£. (A.14)

For composite materials (e.g., phononic crystals) the materials properties p, A and p

become functions of space p — p(r), A — A(r) and p — p(r) (following the distribution

of the different materials constituting the composite). We therefore have to resume de¬

riving and include the spatial derivatives for the mechanical parameters |j, t£ and ^ in

Eqn. A.13.

f, = (^+2W+(A +2^+|^ + ^)+A(f> + |^)\ox ox / ox ox2 ox\oy dz / \dydx dzdx/

dp /dux duy_\ (Q2ux d2u,y\ dp /dux du£\ (d2ux d2uz \

1

dy V dy dx J V dy2 dxdy J dz\dz dx / V dz2 dxdz /

grouping for terms containing A

A(^ +|i+|^)+£*(^ +^ + *!ï) »(W.u) (A.16)V dx1 dudx dzdx J dx \ dx du dz / dx

and p

( (d2ux d2uT d2ux\ /d2ux d2u, d2ux\ l

/7'I V dx2 dy2 dz2 ) V dx2 dy2 dz2 ) )

'dp dur dp, dux dp dvx \ /dp dux dp duy dp, du

dx dx dy dy dz dz'

\dx dx dy dx dz dx

d\i

V-(//Vux) + V-(//—). (A.17)

82

Combining the above produces the wave-equation for inhomogeneous elastic media

du ö\i d

p-^ = V-(/zVw,) + V-(/z—) + ~(AV-u) for x% = x,y,z. (A.18)

A.2 Scattering behavior of single rods

The scattering behavior of an object is characterized by the frequency dependent scattered

field. The most relevant information, however, can be summarized in the scattering

cross-section, o{uj). The scattering cross-section gives the scattering strength (i.e. the

probability for a phonon to be scattered) as a function of the frequency. Resonance states

of the rods can easily be identified in a{uS) and the symmetries of such modes can then

be further analyzed in plots of the scattered field.

The calculation of the scattering cross-section of sound waves at cylinders can be

realized in different ways (See Doolittle and Ueberall,d 27, d'2R Faran,d34 Klironomos and

Economoud-66 or Sutter and Steurerd144). Two methods which have been extensively used

are discussed in the following.

At large distances from the scatterer, the scattering cross-section, a can be defined as

the ratio of intensity scattered to any direction other than that of the incident wave

„ikr

kr

*tot = (A0 - u) eIkr + a--=. (A.19)Vr

Thus the full wave-fields of the incident as well as the scattered waves must be deter¬

mined. The physically most appealing method to calculate the scattered wave-field is by

means of the eigenfunctions of a cylindrical scattering potential derived in the following.

All fields can also be obtained by FDTD-simulations, of course.

Analytical determination of the scattering cross-section of solid

cylinders in a liquid host

First the general eigenfunctions for cylindrical potentials shall be given. The different

fields of the scattering process are then expanded in terms of these functions and by means

of the boundary conditions all fields are then expressed as functions of the amplitude of

the incident field.

Cylindrical Waves

The general wave equation for all types of waves

1 &

(v-?&)*-<> <A-2°>

83

in cylindrical coordinates r —> y/'x2 + y2, Ö —> arctan(y/x), z —> z and

r dr or r2 c>#2 9z

becomes for dip/dz — 0 (2D situation)

'10,0, Id2 1 d2\f--(r-)

\r dr dr+ ?m-7?w>)*=0- (A'22)

canthenbeobtainedusinga

sep

Thereby 6(#) solves

r2dd2 c2dt2,

The wave function ipcanthenbeobtainedusinga separation ansatz V; — B(r)Q(-d)eluJt.

and R(r) solves Bessels differential equation

q2a

^ + m20(t?) = 0 - 0(ï9) - cos{m'd) (A.23)

Ö2i? 1 <9i? m2

The solution for cylindrical wave equation is thus of the form

ipm(iu, t, r, 0) = J^krjcosirniïy^K (A.25)

It can be shown, that these solutions form a full basis set. Any function can thus be

expanded in series of these solutions as

ro

*(w, t, r, tf) = fi*"* ^ cr(t/n(fcr)co*(^)- (A.26)

Field expansions and boundary conditions

The total wave field, \Ptot, can be split into an incident wave, \&int;, a wave inside of

the cylinder, *cyl, and the scattered field, *s. All these fields are expanded in series

of eigenfunctions as shown in Eqn. A.26 (the field inside of the cylinder has to consider

vector fields because of the shear waves present in the solid phase and the expansion must

therefore contain also vector functions).

At the surface of the cylinder, the normal component of *cyi has to match \I>inc + Ws

in equilibrium and additionally the shear stress caused by \&ryi must vanish there. For a

limited degree of expansion, these boundary conditions provide a set of linear equations.

Solving these allows to express the expansion coefficients c^1, csm as functions of c|"c, the

incident field, which can be specified at will. The scattering cross-section can then be

obtained from1 /"27r

^H = ^-r/ \*s(u,<p)\d<p, (A.27)

with I0 the intensity of the incident wave.

84

3

(A

S(A

8>

(b)

800 1200

f [kHz]

*oE

(d)

1

08

06

04

02

0

4000

Figure A 1: Scattering cross-section of a polymeric rod in water (a) and a steel rod in

water (c). The solid black line represents the intensity of the scattered held. In dotted

line, the field scattered by a rigid iod is shown and the led (gray) solid line is the intensity

of the difference field. In (b) the curves obtained from FDTD simulation [red (gray)] are

compared to the analytic results. Harmonic spectrum (d) of the first steel rod resonance

(c).

FDTD-simulation of single rod scattering

In an FDTD simulation a plane-wave pulse is launched towards a rod and the scattering

field is calculated numerically. Thus, the total field, \&tot, is known at all times at all

positions. After subtraction of the incident field, whose unperturbed propagation can

efficiently be calculated in parallel, the ratio a can be determined by integration of the in¬

tensities measured on a circular detector (at large r) around the seatterer and normalizing

this intensity with the one of the incident plane-wave

CT^) = 2^Ät//_ l*tot(w,f,v)-*in(a;,t,^)|d¥?(// (A.28)

A comparison of the analytically calculated and the simulated cross-sections in Fig. A.l

reveals a very good agreement at low frequencies. The frequencies of the resonance1 peaks

agrees exactly. On the other hand, at higher frequencies, convergence problems occur

(even for resolutions of dx — r/1000) For the hard scatterers this was not a problem

because the frequency range of interest was low with respect to ur/c^. Further contribu¬

tions to the deviation of the curves are caused by differences in the setup (i.e. boundaries,

overestimated tymc in the forward direction in the shadow of the seatterer, distance of

detector, etc.).

85

400 800

f [kHz]

1200

„08

*».Ora 04

(b)

-

m=2' / X

/ -

"

m=1.

i

y

\/

/ '* m=3

400 800

f [kHz]

Figure A.2: Contributions of the first three components <^nof cylindrical waves to the

total scattered fields for polymeric rods (a) and steel rods (b) in water.

Resonant vs. point-like scatterers

In this section we want to briefly discuss two prominent cases of scattering types. Soft

polymeric rods in water scatter according to the cross-section shown in Fig. A.1(a). In the

frequency range of interest (wa0/cmatnx ~ 1), there are many resonances. Their scattering

strength is significantly higher than at frequencies between two successive resonances.

Also the symmetires of the scattered fields changes completely from one resonance to the

next. It is therefore not surprising, that the resonance states dominate the formation of

the band structure of phononic crystals.

The scattering cross-section of a steel rod in water is shown in Fig. A.l(c). The

first resonance peak appears at ~ 2.4 MHz. In all the range below, the curve is almost

identical with the one of a ideal rigid scattcrer and almost 'ndependent of the frequency.

The symmetry of the scattered field changes smoothly with the frequency because the

contributions of the different cylindrical harmonics is shown in Fig. A.2(b) to replace

each other smoothly. In other words, up to 2 MHz the scattered fields do not carry much

information about the nature of these scattering objects. Thus, the formation of the band

structure is governed primarily by the type of arrangement of the scatterers.

A.3 Transmittance and attenuation regimes

Other than in real solid-state physics in PNC's the exact structure is always known. In

some characteristic cases, this allows simple estimations of the form of wave-functions

involved and therewith of the type of transmission or attenuation spectra.

Regular Bloch bands

In periodic structures the most general types of wave-functions are the Bloch-waves. These

functions are of the form

$(r) = C/(r)e'kV^, with U(t) = f/(r + T) (A.29)

86

where T is a primitive lattice vector of the structure. For every k the corresponding

frequency u(k) is an eigenvalue of the corresponding Hamilton operator. These; extended

Bloch waves (extended over the whole PNC) can propagate unhindered (unattenuated)

through the PNC from one end to the other. In finite PNC's, surfaces effects of course

affect the transmission spectra. The most prominent of these, the Fabry-Perot resonances,

stem from multiple surface reflections and induce a modulation with the period A/ =

ceff/2d to the low frequency range of transmittance (with the sample; thickness d). Of

course, if some of the materials employed are absorbing the Bloch-waves are damped and

transmission gently decreases towards higher frequencies.

In electronic structures, usually only solutions of the; wave equation in ranges of the

bands are considered (i.e. waves with real wave vectors). If k is allowed to be complex, on

the other hand, there is always a solution of the wave-equation (or more precisely, there

are always two,d5ü'd119) even in ranges of the bandgaps. There k — zq, with q G R3 is

purely imaginary and the corresponding waves,

#(r) = [/(r^V"', (A.30)

are obviously exponentially decaying with r and are therefore called evanescent (the wave

with exp(+kr) occurs for bandgaps opening at k = 0). Such waves are localized to the

surfaces of a PNC. In the complex k-space, the bands form continuous lines, which do

not stop abruptly at the edge of a regular Bloch band. For the evanescent waves the

imaginary wave-vectors act like attenuation parameters and the logarithmic transmission

ratio becomes directly proportional to the complex band function log(T) ~ —Im[k(uj)]

d, with d the sample thickness.'1118 In infinite PNC's, of course such waves are not of

relevance.

Diffusive wave propagation

The complete opposite of the periodically ordered PNC is the random arrangement of

scatterers. In case of strongly disordered systems and under absence of strong resonances

of the scatterers, wave propagation becomes diffusive in nature.dq9 The energy density of

the waves, $, then obeys the diffusion equation

(— + DV2)$ = 0 (A.31)

with D being the diffusion constant D. Prerequisites for the applicability of this simplified

picture is complete loss of coherence, which is easily fulfilled if scatterers are arranged

in an uncorrelated manner. Elements of the theory which are being used hen1 are very

fundamental (see for instance Gerthscn and Vogeld'4ü) and can be illustrated by the picture

of a particle P with radius r moving through a system of other particles P% with radius rx

and number density n. If these an; stationary, the mobile particle incurs a collision when

87

0 1000 2000 3000 4000

f [kHz]

Figure A.3: Hypothetic evolution of transmission with the sample size according to a

diffusive wave propagation regime (black). In red (gray) the spectrum of QPNC-14 is

shown (steel/water).

it gets closer to a stationary particle than the cross section ac — n(r + rt)2. Along the

particle's path, x, a cylinder of volume a • x contains all a • n x particles which provoke

scattering. The path length between two successive collisions is on average xc = lmip —

1/na. Now, in case of N incident particles in a beam, the probability for a collision is

NdP = Nandx = dNc, which leads to dN/dx — —anN when scattered particles are

considered as loss. The number of particles of the beam reduces as

N(x) = T{x) = e-'7nx^e~r^, (A.32)

Diffusive wave propagation can be described essentially by the mean free path. An esti¬

mation for /mfp can be obtained by weighting the scattering cross section of the rods (see

section A.2) with the reduction of the forward component of the scattered wave intensity

i r-—— = n / a(uj, 0)(1 - cosö)d0 (A.33)'mfp(^) ./O

with n being the number density of scattorers. On the other hand, condition A.32 allows a

simple guess of /mfp by studying T vs. sample size. In experiment, though, this approach

proves to be rather difficult because of the statistics involved. Transmission measurements

usually use intensities averaged over a large area of the sample surface. This is necessary

to get better signal to noise ratios. While for periodic systems this is not a problem,

because the Bloch modes are extended over the whole crystal, in non-periodic systems

the averaging accounts for different scattering paths which do not necessarily cohere. The

intensity of waves leaving the sample at different locations on the surface; produce a speckle;

pattern of fluctuating intensity. In order to correctly average each wave path, a spatially

resolved measurement is necessary.dl0°

In Fig. A.3 a transmission curve of the QPNC-14 is compared to a hypothetical evolu¬

tion of transmission with the sample size if perfectly diffusive wave propagation is assumed.

There is a clear plateau of average transmission in the range above 2 MHz (interrupted

only by some weak resonances). In the stronger scattering range; betaw 2 MHz the curves

88

are fai more deviating. Transmittance of waves with these frequencies clearly reflect

structural features of the quasiperiodic arrangement, which cause, for instance, the sharp

edge at — 300 kHz. This edge wc have identified as the lower edge of the first bandgap of

the PNC with the periodic average structure of the 14-fold tiling. Clearly the comparison

demonstrates the inadequacy of diffusive propagation regime for the example at hand.

Localization and renormalized diffusion

In real PNC's, neither peifect periodicity nor real randomness can ever be achieved.

There are always two competing forces at work. Order on the one hand, which drives

for extended wave1-functions and disorder on the other, which drives towaids complete

incoherence of wave propagation (i.e. diffusion). Intermediate between these two extreme

regimes, gradually increasing localization occurs. Localization is a process happening at

the edge of coherent and incoherent wave propagation but still is an interfeience effect.

Localization due to disorder has first been addressed by Anderson in 1958.d7 His

seminal input spurred a lot of work on this specific diagonally disordered Hamiltonians

(diagonal elements of the hamiltonian (on-site energies) deviate1 from an average energy by

a certain AE). Without lequirement of a specific type of structure all waves can be shown

to localize if this disorder, AE, exceeds a certain threshold value (Anderson transition).

Anderson localization is to be distinguished from localized modes due to resonances.

We want to approach localization in two steps. A first step is given by the Ioffe-

Regel criterion k • lmfp ~ l.d29 If the scattering strength of scatterers is large and their

arrangement is disordered and not too diluted (a strongly scattering medium), the mean

free path length of wave propagation becomes short, shorter than the wave length even.

If this is the case, the uncertainity of k is getting large. The spatial periodic nature

of a wave is lost thereby. Or, if Ak is getting large (spiead in reciprocal space), the

'wave' is becoming confined in direct space. The wave is not localized thereby, it is still

propagating. Only the1 coherence of its propagation is lost to a large extent.

In literature the onset of localization is usually assigned to the enhancement of coherent

backscattering/1 5> d 123 This effect is due to the interaction of two waves traveling on

exactly the same paths through the medium just in opposite directions. A significant

increase of the intensity scattered in the backwaid duection results. Thus, the overall

transmission is reduced thereby, which is termed weak localization.

With gradually stronger localization transmittance is reduced more and more until

complete isolation is achieved (i.e. conductivity, diffusion constant, etc. —> 0). Trans¬

mittance, howevei, cannot easily be guessed as in the case of classical diffusion. It is a

typical feature of systems with localized waves that properties related to conductivities

are depending on the size of the system as long as their spatial extensions are small with

respect to the localization length £. This is so, because the exponentially decaying tails

of the wave functions can still feel the boundaries and absorb or release energy from the

89

exterior. In such small systems, properties like conductivities are even depending on the

system size.dl2:i Regarding the strength of localization, two fundamental aspects must be

discussed. According to Shcngd123 fundamental difference must be made between disor¬

dered periodic systems and purely random systems. Periodic systems feature a special

type of waves, close to a band edge, whose velocity is, by means of dispersion, already

close to zero. Such waves can be expected to first localize once disorder is introduced.

With gradually stronger disorder AE, more and more waves towards the center of a band

fall for localization. At AEclü, the Anderson transition, this mobility edge reaches the

band center and all waves are localized. In real random systems there; is no such aid for

localization. Therefore AEcrit is larger.

The second most crucial parameter is the dimensionality of the system. In ID and 2D

Anderson Hamiltonians any degree of disorder AE suffices to create some localized states

(see, for instance, Economoud31). In 3D this is not the case. A first threshold value for

AE is required for the first states to localize.

For 2D random arrangements of hard scatterers, Condat and Kirkpatrkkd-22 have

calculated the localization length £(u). At 0.35 filling fraction of scatterers, they claim

a minimal length of £/a ~ 100 if a is the size of the scatterer. Since at edges of bands in

periodic or other ordered arrangements localization is much easier established, localization

could be expected also in samples of the size analyzed in this work. In 3D PNC's the

localization behavior has been theoretically studied by Sainidou and co-workers.dii8 They

found that cubic PNC's with displacive disorder are weakly and such with substitutional

disorder strongly subject to localization (periodic but 3D). In both systems a severe

enlargement of the width of bandgaps was observed.

In order to establish a connection with the description of the diffusive wave propagation

in the previous section, a short overview of the formalism of Condat and Kirkpatrickd*22' d"23

as well as Shengd123 shall be given. They state their theory in the framework of the self-

consistent diagrammatic approach (SCDA) to localization. The system under study is

a multiple scattering process in a random arrangement of scattering bodies. The phys¬

ical quantity of interest is thereby the evolution of the elastic energy density e:(r, t) =

Po[V$(r,t)]2. On a macroscopic level (for long times scales and large distances) this evo¬

lution is diffusion-like, which is in agreement with the conservation of energy (in absence

of absorption). Its propagator P is given by

P(r,£|r',0)-G2M|r',0)

and the diffusion equation Eqn. A.31

(Z)V2 + jjP(r,t\r', 0) - ö(r - r')ô(t). (A.35)

Thereby, G is the Green's function of the system from which the wave functions can be

(A.34)

90

obtained by

W(M) = y"G(r,f|r',0)/(r>/r', (A.36)

and which itself solves the wave equation Eqn. A. 18 as well as appropriate boundary

conditions [/(r') is then the excitation of the system, e.g., an incident pulse]. From the

Laplace transform of P

1 f°°

PMW) = ~J G%+(r\v')GE_(r'\r)dE, (A.37)

with/OO

G+(r|r')= / e«E±a)tG{r,t\r',Q)dt (A.38)Jo

and E± = E ± | it becomes clear, that P is a two body propagator. This may seem a

formal affair, but its physical implications are vital. The averaging of [Pfc>(r|r')]average

over a homogeneous random arrangement of scatterers is not performed on a single par¬

ticle's propagator (i.e. the temporal displacement of a particle, a random walk) but on a

property which intrinsically includes interaction of waves (an interference effect as initially

mentioned). The Fourier transform of [PB,u;(r|r')]average then becomes for k,u —> 0

[P«AE)U^ ~

_lu + D{E^)^(A-39)

P has a diffusive pole, which designates correct forms of e(r, t). Thereby, D is a frequency

dependent Diffusion constant. Diffusion is understood here in a sense, that at distances

far away over long time scales the distribution of scattered intensity would follow the laws

of diffusion, while locally full multiple scattering processes are at work.

The evaluation of the diffusion constant can now account for the typical localization

effects. The diffusion constant D is reduced by an amount, which is partially ascribed to

the coherent backscattering effect. The coherent interaction of two waves, which travel

exactly on the same path through the sample only in opposite directions results in a severe

enhancement of the backscattered intensity. The forward propagation of the average

intensity is reduced thereby - this is equivalent with a renormalization of the diffusion

constant.

91

Appendix B

Computational methods used

The methods used through out the thesis project aim to simply solve the fundamental

Eqn. A. 18 discussed in Appendix A. The approaches are very different and each of the

methods has its very own advantages. We give here only short sketches of these methods

and refer the reader to more extensive source. Descriptions and comparisons of these as

well as other methods used in the field are given, for instance by Sigalas et a/.dJ30 and

Soukoulis.d-133

B.l The plane wave expansion method

The plain wave expansion method (PWE) is transforming all variables of the wave; equa¬

tion into Fourier space. In Fourier space they are connected by systems of linear equations

which can be easily solved numerically.

The PWE-method was the first of the classical band structure calculation schemes

applied to phononic crystals.*169 In this procedure the periodic variations of the mechanical

properties p and cl are exploited to express them as Fourier scries

P(r) - p(r) = ]Tp(GmyG~rGm

c,(r) -> cl(r) = ^cf(GmyG-r. (B.l)

Due to the; periodicity of the system, the wave functions solving the wave equation must

be of Bloch type

u(r,G) = C/(r)VGr (B.2)

with U(t) — U(r + T) (Ta lattice; vector of the structure). Thereby a discretization

(required for computation) is achieved in Fourier space. Substitution of Eqns.B.l,B.2

into Eqn. A. 18 gives

p(Gm)|^(G) = y • (/i(Gm)Vu,(G)) + V • ^(GJ^pi) + £ (A(Gm)V u(G))(B.3)

92

for Xi = x,y,z. Since the Fourier series Eqns.B.l are known the wave function u can be

obtained. Yet, Eqn. B.3 is an infinite set of equations which relate the Fourier coefficients

of the wave function u(G) with those of the mechanical properties F(Gm). Neglecting

coefficients for wave vectors larger than kmax provides a good approximation and the

problem reduces to solving N = kmax/Ak linear equations u(G) — /(Gm) (i.e. finding

the eigenvalue of an N x TV matrix).

Specific descriptions of this formalism can be found in several of the early articles on

phononic crystals. A rather comprehensive one is given by Vasseur and co-workers.d151

For the electronic case basic text books explain the underlying physics.'1'8

B.2 The finite difference time domain approximation

The finite difference time; domain method (FDTD) is a very powerful but basically rather

unphysical approach. The wave equation (or any other linear partial differential equation)

can be solved by approximating the derivatives of the wave function ù and Vru by central

finite differences

..

,u(r, t + dt) — u(r, t — dt)

u(r,t) ->->

'- and

d_ u{x + dx,t)-u(x-dx,t)dxU[ ' j

2dx[ '

The FDTD method allows to calculate the propagation of waves in a limited spatial

domain, with well defined boundaries, over a certain period of time. In this domain, the

wave field is discretized in space and time with suitably small Ax and At. Practically

this can be done by storing the displacement u as well as the field of tensions t in arrays

(?' — l,imax;i — It jmax) °f a computer program. Starting with an initial distribution of

displacements and tensions the evolution of these fields to a future state can be obtained

as follows. The tension a node (?', j), t(z • Ax, j Ay, k At), can be calculated from the

displacements of the surrounding grid nodes. This resulting tension can exert a force

on this grid point and accelerate it for a time Ai. The displacement at a node (i,j),

u(z • Ax, j • Ay, k At), thus follows from the displacement as well as the tension at the

node at a previous time point. From this field, again, the field of tensions can be updated

etc..

The discretization

Generally the resolution Ax must be small compared with typical sizes of objects or their

features (i.e., small with respect to the shortest wave length that may occur). Higher

resolution in space is thereby on the expense of computation time. One notable excep¬

tion is the staggered grid approach by Yee.d'166 In this type of meshing the displace¬

ment field and the field of tensions are discretized on two different grids. These grids

93

are shifted one against the other along half the diagonal of a grid cell. The tension

at grid node (i + 1/2, j + l/2)u, tî+1/2j+i/2, is then obtained from the displacements

Uj^Ui+i^Ujj+^Uj+^+x. The displacement field is thereby linearized only over one cell

instead of two. The accuracy of the resulting tension is increased without finer meshing.

The displacement field after At is then obtained from the tensions at surrounding nodes

v,_i/2j-i/2, vt_1/2+i,,+i/2, v,+1/2j-i/2, v,+i/2,j+i/2. While in electrodynamics also the tem¬

poral discretization of the two fields is shifted by At/2 for elastic waves they are kept the

same/126' di30

In principle, the discretization of space, Ax, and time, At can be chosen arbitrarily.

However, numerically stable results can be guaranteed only if At is smaller than the magic

time step

At <,

X(B.5)

~

2cVAx"2 + Ay~2

Boundary Conditions

The evolution of the wave field can only be computed for a well defined domain. For

nodes at the boundaries of this domain (e.g., ti/2,.,+1/2) the above described procedure of

calculating the future fields from the fields at surrounding cells cannot work because not

all the cells exist.

Periodic boundary conditions are introduced if instead of the values of the missing

cells, e.g., Uo,;, the values of cells on the other end of the domain, uïmaxJ, are used to

update ti/2,^+1/2- The; by the ends of the domain can be directly linked.

If an open domain is modeled, a wave exiting the domain at one end is not expected

to enter it again on another, neither should it be reflected from hard Dirichlet boundaries.

Absorbing boundary conditions (ABC) are then required. Currently, two types of ABSs

are being used. In the perfectly matched layer technique a layer of absorbing material is

attached to the domain in such a way that no impedance contrast results at the interface.

The waves are then damped in the absorbing layer. In the differential type of ABCs, as

are the Mur ABCs,d94 a one-way wave propagator acts onto the field at the boundaries.

This operator forces the field to solve the wave equation of an outgoing wave and this in

return can be shown to be sufficient to guarantee a suppression of reflected waves.d32 For

an overview of ABCs see Taflove and Hagness.d1'17

B.3 The multiple scattering method

The multiple scattering technique; (MST) has been used only brevely to accompany the

experiment described in Sec. 2.4.

The fundamental physics of the approach can be found in many standard text books

in treatments of the KKR-method for electronic band structure calculations.118'089 For

94

a specific description of the program used the reader is referred to the author's article,

Sainidou and co-workers.d116, d117

The fundamental MST equations describe the total wave field VKr) as the snm OI the

initially undisturbed field, t/;o, and the fields scattered from all individual scatterers, Vf°

V;(r) = ^o(r) + J>r(r) (B.6)

On the other hand, the field incident to scatterer i, ij)c, consists of the undisturbed field

and the sum of fields scattered from all other scatterers

0r(r)=V;o(r) + J>r(r)- (B-7)

The sum V;,mc(r) + V;fc(r) readily gives the full wave field again. The dependence of

0fc on V;J"r constitutes the problem of single object scattering and can be solved, for

instance by help of the boundary conditions at the matrix scatterer interface (see appendix

A.2). The field scattered by a scatterer centered at 1 can be generally described by a

scattering potential v(lj, r' — 1). The intensity of the scattered wave can be assumed to be

proportional to the incident field amplitude (v is essentially a scattering cross section).

The influence of local scattering due to v(<^, r' - 1) at r' on the whole field can be well

described by a Green's function G(u>, r — r'). The Green's function can transfer the effects

of the local excitations at r' to any other site r as

V;(r) - / G(u, r - r')v(Lj, r' - \)^(r')dr'. (B.8)

If there is a whole set of scatterers the total scattered field is again just the sum of waves

scattered from all lattice nodes

V;(r) = Y1 [ G(w' r - r')v(w' r' " WOdr'. (B.9)l

'

For periodic arrangements of scatterers the Bloch theorem can be easily included into

the last equation, which can then be solved using the variational principle. Eqn. B.9

clearly illustrates that the problem of scattering at single objects can be separated from

the structural aspects of the problem. This is the beauty of the method and also the

reason for its efficiency.

95

Appendix C

Systematic description of quasiperiodicstructures

In this appendix quasiperiodic structures and especially their higher-dimensional descrip¬

tion shall be discussed in more detail and in more illustrative ways.

A systematic description and categorization of quasiperiodic structures (quasilattices)

first has to distinguish structures, which are quasiperiodic in one-, two- or three dimensions

(and periodic 01 constant in the other ones). In the course of the cm rent thesis all

types have been treated but the focus clearly is on planar quasilattices. In 2D and 3D,

quasilattices are further specified by the degree of rotational symmetry of their diffraction

pattems. In ID structures the recursive generation (i.e. their degree of scaling symmetry)

is chaiacteristic. How and why exactly this can be achieved is described in this section

as well as in several reviews and books.d 7R< d 131' d i32' d 136' d lf>4

Crystallographic prerequisites

It is well known, that 2D (3D) translational periodic structures can exhibit 2-, 3-, 4-

and 6-fold rotational symmetries only. A 5-fold axis, or othei elements of the non-

crystallographic point groups (including rotational symmetries of any degree a e N other

than the above mentioned) are prohibited. They are prohibited because tiles (e.g., regular

pentagons) which have such a symmetry (e.g., a 5-fold axis) cannot be densely packed to

cover the plane (space) without leaving voids.

If periodicity is dropped as a restriction though, there is no need anymore foi a single

unit cell, but arrangements of different unit tiles are possible. Two kinds of rhombs, for

instance, with angles Qi — 2-7r/5 and a2 — tt/5 can be arranged to densely cover the

plane in such a way that even 10-fold global symmetry results. Quasiperiodic structures

are special cases of such airangemcnts of several constituent prototiles and have the four

properties mentioned in the1 introductory chapter on structures in Sec. 1.2

96

• perfect short- and long-range order

• absence of a translational period (i.e. absence; of a unit cell)

• a Fourier spectrum consisting of a dense set of singular <5-peaks

• Scaling symmetry of the Fourier spectrum.

In ID, there are no such symmetry issues (or packing problems) by which quasiperiodic

and periodic chains could be distinguished. Nevertheless, the four points equally apply

to them.

Higher-dimensional description

The higher-dimensional description introduced in this section is very helpful for the sys-

tematics of quasiperiodic structures because it provides a simple link between a structure

and its Fourier transform, in which the classifying symmetry is defined/154, d136

Besides the systematica, the higher-dimensional description is used by algorithms for

the generation of tilings, the cut and project method (or the closely related strip projection

method). This procedure is illustrated in the following ID example. Other methods

frequently used are the dual-grid method'14' d-3, d95 and methods using inflation/deflation

step-growth of tilings (see Sec. 2.2 on pinwheel tilings).

A ID quasiperiodic sequence from a 2D square lattice

This example is intended to provide a simple recipe for the construction of a quasiperiodic

structure. Some of the initial assumptions are explained and justified only at the end.

Given a 2D square lattice spanned by the d basis (di,d2)v = {(1,T), (T, _i)}v with

t the golden ratio. V denotes the Cartesian reference space consisting of V" || ei the

physical parallel space and V1 || e2 the perpendicular space. The slopes of both basis

vectors d; with respect to V (r/1 and — 1/r) are both irrational. As a consequence, no

lattice point other than the origin intersects one of either subspaces, V" or V1. Then,

from this square lattice certain points are selected by a strip S around V" of a certain

width. Orthogonal projection of these selected points onto physical space V" produces a

sequence of points with only two different distances S — di ei and L — d2 • ei (or r±n

scaled, depending on the strip width). As a consequence of the fact, that no lattice point

other than the origin is lying in VH the sequence contains no section equivalent to that

around the origin, it does not repeat itself and is thus not periodic.

Taking then the modulus of all vertices of the sequence within a square lattice unit

cell produces lines centered on the lattice nodes [see Fig. C.l(a)]. Knowledge of these

lines, termed the atomic surfaces, allows construction of the sequence in an alternative

way. The points of intersection of these atomic surfaces with V" select the points of the

97

Figure Cl: Construction of the; Fibonacci sequence from a 2D square lattice (a). The

points of intersection of the parallel space V'l with the atomic surfaces are the vertices of

the sequence. The Diffraction pattern of the sequence (b) consists of the projection of all

vertices of the reciprocal square lattice onto H" with a weighting according to their Hx

component according to ^(H-1) (c).

square lattice, which constitute the sequence. These hyper-objects are nothing else but

domains of perpendicular space which select the tiling-vertices from an infinite point-set.

Their extension in V" must be zero, because they have no physical meaning but are

purely part of an artificial construction scheme. Nevertheless, the unit cell of the square

lattice together with the atomic surfaces can be understc jd as a unit cell for the infinite

quasiperiodic sequence. The aperiodic ID structure is periodic in 2D, which is the crucial

advantage of higher-dimensional approach.

The square reciprocal lattice is spanned by (d^d^v = 1/(1 + r2) (di,d2)v- The

Fourier transform of the convolution of the real square lattice with the selecting atomic

surface corresponds to a multiplication of the reciprocal lattice with the Fourier trans¬

form of an atomic surface, which is formally termed geometrical form factor gfc(Hx) [see

Fig. C.l(c)]. The Fourier transform of the intersection of the real lattice, decorated with

atomic surfaces, with V^ (i.e., a multiplication with a £(V" - 0) function) corresponds

to a projection of the whole pfc(Hx)-weighted reciprocal lattice onto H'1. This produces

the ID Fourier transform of the sequence, which comprises an infinite number of Bragg

peaks. Reciprocal lattice points with large components in H-1 generally produce weak

reflections and those with small such components produce stronger reflections, according

to ^(H-1). Each reflection can be identified by two proper indices (its coordinates in the

d* basis). Again as a consequence of the irrational slopes of (d^d^v with respect to H",

no two reflections project onto the same point on H" but instead, they densely cover it.

In summary, a sequence of two segments is formed, which does not have any period

and whose Fourier spectrum consists of a dense set of singular diffraction peaks. We thus

98

deduce, the structure must be quasiperiodic. And indeed the chosen example sequence is

the well known Fibonacci chain (for certain widths of S).

The remaining questions concerning the description of our Fibonacci example are, how

can we know in advance; which lattice and what widths of the selection strip must be taken.

First, the basis vectors of the square lattice are the eigenvectors of the recursive growth

matrix of the Fibonacci sequence (i.e. eigenvectors of its scaling symmetry operator)

(S, L) -, (S, L) f J 1\=(L,L + S)=t-(S,L),

with L and S the long and short elements of the sequence. Or the other way round, the

sequence; is lying in a subspace of a 2D Cartesian lattice. This whole subspace is invariant

under scaling with factor r, and so is the sequence itself. The width of the strip is not

clear in advance. Yet, variation of the width of the strip reveals, that the same sequences

appear again and again at certain values. These different (singular) values for the strip

width producing exactly a Fibonacci chain are reflecting again its self-similarity. The

atomic surfaces can either be determined subsequently, when the singular values for the

strip width are known or it can be deduce from the; closeness condition. This condition

guarantees that from every unit cell of the square lattice; only (maximally) one point is

selected for the sequence [see the dashed horizontal line connecting two atomic surfaces

in the first unit cell of the square lattice in Fig. C.l(a)]. Their length is then equivalent

with the projection of a unit cell of the square lattice onto V-1.

A 2D quasiperiodic structures from higher dimensional lattices

In order to obtain a structure, which is quasiperiodic in two (three) dimensions, the rank

n of the; Z-module of its embedding space must certainly be large;r than two (thre;e). Its

specific dimension is given by the desired symmetry of the structure. The construction

of a tiling of, for instance, 10-fold symmetry can certainly be; achie'ved by using a 10D

cubic lattice. In a first step the cigenspaces of a 10-fold axis (i.e., subspaces, which remain

invariant under cyclic permutation of the basis vectors di,..., d10) are determined. The

2D tiling must lie in such an eigenspace; in order to have a 10-fold symmetry. This

eigenspace is then assigned V" while all other dimensions constitute V1, which is of

n — 2 — 8 dimensions. A transformation of coe)rdinate,s with V^1 brings everything

into the form analogous to that in the Fibonacci example in the previous subsection

(a Cartesian reference lattice and a skewly embedded e;ubie- d basis). A quasiperiodic

structure can be obtained by projection of vertices of the Z10-module selected by a strip

S\\ V" ontoVH.

A structure featuring 10-fold diffraction symmetry, however, can be obtained also from

five or even four dimensions (because of the centrosymmetry of all diffraction patterns).

This is a consequence of the fact that the projection e)f the; 5-fold reciproe-al lattice onto

99

V" produces also a 10-fold symmetric pattern. This is analogous to the example of a 3D

cubic lattice projected along the [1 1 1] direction onto a (1 1 1) plane. The projection

along [1 1 1] transforms the 3 axis into a 6-fold axis. This effect N —> 2N applies to all

odd N. Further reductions of the required rank of the Z-module can be achieved by using

non-cubic d basis, which better fit the desired symmetry (for instance, the dodecagonal

structure can be produced from 12, 6 and 4 dimensions). The minimal dimension required

is given by Euler's totient function <f>(n) (i.e., the number of integers smaller or equal to

n with no common devisor greater than ldn).

The atomic surfaces in higher dimensions have a dimension smaller than or equal to

the dimension of V1- and may adopt very complex shapes or also consist of several uncon¬

nected pieces. The closeness condition equally applies and becomes rather complicated

with the complicated shapes of the atomic surfaces.

Quasiperiodic structures can be conveniently and systematically described by higher-

dimensional periodic unit cells, given by a basis and the exact shape of the atomic surfaces

therein. Additionally a direct link to the Fourier transform of the structure is established.

Periodic average structures

In the introduction to this section we already stated that there is a certain degree of peri¬

odicity associated with every quasiperiodic structure. This degree varies for the different

types of structures and can be used to characterize them (see article 5 in Sec. 2.3). The

periodic average strucjire (PAS) of a quasiperiodic tiling is defined as a periodic structure

whose reciprocal lattice basis vectors point to the strongest reflections in the diffraction

image of the tiling.018' dl3T The concept of the PAS has been already successfully applied

to surface science of quasicrystals.dl38> dlfi1Deposition of crystalline matter on quasicrys-

tal surfaces was found to evolve in such a way that coincidence of the crystal lattice with

the PAS is best achieved.

In the higher-dimensional description, the PAS of a 2D tiling can be understood as

follows: choosing two strong reflections, i.e. two reciprocal lattice vectors from the infinite

Z-module, corresponds to a cut of a plane through the nD reciprocal embedding space.

The plane spanned by the origin and the two reflections contains a subset of reflections.

For large n it is easier to imagine the cutting as stepwise reduction of dimensions. A nD

polytope is cut by a (n — 1)D hyperplane to give a (n — 1)D polytope. This polytope

is cut by a (n - 2)D hyperplane to give a (n - 2)D polytope etc.. Since the 2D plane

used in the last step must cut an infinite plane out of the original reciprocal lattice,

all previous cuts must cut parallel to this plane (include this plane). To give a simple

example, suppose we want to choose one reflection additional to the origin from a 3D

reciprocal lattice. We can then first cut space with any arbitrary plane including the

two reflections and then cut the line out of this plane. In direct space; this is equivalent

100

Figure C.2: Construction of the periodic average structure of the Fibonacci sequence

from in the 2D square lattice (a). The oblique projection of the unit cell of the square;

lattice along the (l,-l)-direction results in a periodic sequence [aPAs — (3 — t)/S\ and

in projected atomic surfaces with a non-zero extension in V" (rPAs)- The direction of

projection is perpendicular to the line (fixed by the (0 0) and the (1 1) reflections) cutting

the reciprocal space (b).

(again due to the analogy of projecting/cutting in real and cutting/projecting in Fourier

space) with multiple projections along n — 2 directions which are all perpendicular to the

chosen reciprocal plane and, because they have to cause a reduction of dimensions, are

all linearly independent. Just as the specific cuts in Fourier space are arbitrary as long as

they produce in the end the chosen reciprocal plane, so a.j the directions of projection, as

long as they are all perpendicular to the chosen reciprocal plane and linearly independent.

Now, because these directions are oblique to the physical parallel space the projections

elevate the incommensurability and a periodic structure results (see the procedure for the

ID Fibonacci case in Fig. C.2).

The procedure yields a periodic lattice, specified by its two main reflections in recip¬

rocal space. This lattice corresponds to the projection of the nD lattice onto the physical

parallel space. Because the atomic surfaces are thereby projected accordingly, they obtain

a non-zero extension in V" and become visible, usually in a distorted way. Additionally,

for most directions of projection the projected atomic surfaces on neighboring PAS ver¬

tices overlap and cover all of VK For good PAS, however, they may be well separated and

rather small. These projections of the atomic surfaces indicate how far the vertices of the

quasilattice can deviate from its periodic average; structure (see Fig. C.3). Their width

can thus be interpreted as a degree of aperiodicity. In simple cases like the Fibonacci

sequence the atomic surfaces consist of only one line or in simple tilings like the octagonal

oik;, of a simple octagon. They can be visualized either by taking the modulus of the tiling

vertex coordinates within the PAS unit cell [as in Fig. C.3(b)J or by actually projecting

the atomic surfaces from the higher-dimensional embedding space. For more complex

101

Figure C.3: The periodic average structure of an octagonal tiling (a) defined by the (Olli)and (110Ï) reflections consists of a square lattice hosting regular octagons. The deviation

of the quasilattice vertices from this PAS is only small. Tin; modulus of the tiling vertices

within the PAS unit cell shows the homogeneous occupation of the projected atomic

surface (b).

tilings the modulus procedure is not sufficient to get a clear picture of the projected faces

because different sections of it may overlap. In such cases the convex hulls of all pieces of

the atomic surface must be calculated and projected to parallel space.

For the Fibonacci sequence there is a one-to-one correspondence of vertices of the

sequence and the PAS (i.e. (-very projected atomic surface on the PAS is hosting a vertex of

the sequence). The structure; can be described by a quasiperiodic modulation of a periodic

sequence. In 2D and 3D, this is not generally the case. As can be seen in Fig. C.3(a) for

some of the projected atomic surfaces there is no associated tiling vertex. The occupancy

ratio is reflecting the difference in vertex densities between the quasiperiodic tiling and

its PAS. Quasiperiodic structures are therefore not commonly considered as modulated

structures.

102

Bibliography

[d.l] http://phys.lsu.edu/ jdowling/pbgbib.html1

fd.2| http://www.phys.uoa.gr/pliononicR/

[d.3] http://mcs.open.ac.uk/ugg2/quasi.shtml

[d.4] J.L. Aragon, G.G. Naumis, and M. Torres, Acta Cryst. A58, 352 (2002).

[d.5] E. Akkermans, P.E. Wolf, R. Maynard, and G. Maret, J. Phys. France 49, 77

(1988).

[d.6l E.L. Albuquerque and M.G. Cottam, Phys. Rep. 376, 225 (2003).

[d.7] P.M. Anderson, Phys. Rev. 109, 1492 (1958).

[d.8] N.W. Ashcroft and N.D. Mermin, Solid State Physics, Thomson Learning, London

(1976).

[d.9] F. Axel and D. Gratias, Beyond Quasicrystals, Springer Verlag, Berlin (1995).

Id.10] H. Aynaou, E.H. El Boudouti, Y.El Hassouani, A.Akjouj, B. Djafari-Rouhani,

J. Vasseur, A. Benomar, and V.R. Velasco, Phys. Rev. E 72, 056601 (2005).

[d.ll] M. Baake in: J.B. Suck, M. Schreiber, and P. Haussier (Ed.), Quasicrystals: In¬

troduction to Structure, Physical Properties, and Applications, 17, Springer Berlin

(2002).

[d.l2j M. Baake, D. Frettlöh, and U. Grimm, Phil. Mag. 87 (in press, 2007).

[d.13] R. Berenson and J.L. Birman, Phys. Rev. B 34, 8926 (1986).

fd.14] M. Bcrtolotti, P. Masculli, and C. Sibilia, Optics Letters 19, 777 (1994).

[d.15] R.T. Bratfalean, A.C. Peacock, N.G.R. Broderick, K. Gallo, and R. Lewen,

Opt. Lett. 30, 424 (2005).

1 Titles of the bibliography appear in alphabetical order of the first author's name.

103

[d.16] S. David, A. Chelnokov, and J.M. Lourtioz, IEEE J. Quant. Electr. 37, 1427

(2001).

d.17] N.G. De Brujn, J. Phys. (Paris) 47, 9 (1986).

d.18] A. Cervellino and W. Steurer, Acta. Cryst. A 58, 180 (2002).

d.19] Y.S. Chan, CT. Chan, and Z.Y. Liu, Phys. Rev. Lett. 80, 956 (1998).

d.20] S.S.M. Cheng, L.M. Li, CT. Chan, and Z.Q. Zhang, Phys. Rev. B 59, 4091

(1999).

d.21] A.N. Cleland, D.R. Schmidt, and CS. Yung, Phys. Rev. B 64, 172301 (2001).

d.22] C.A. Condat and T.R. Kirkpatrick, Phys. Rev. Lett. 58, 226 (1987).

d.23] C.A. Condat and T.R. Kirkpatrick in: P. Sheng (Ed.), Scattering and localization of

classical waves in random media (World Scientific scries on directions in condensed

matter physics ; vol. 8), World Scientific, cop., Singapore (1990).

d.24] L. Dal Negro, M. Stolfi, Y. Yi, J. Michel, X. Duan, L.C Kimerling, J. LeBlanc,

and J. Haavisto, Appl. Phys. Lett. 84, 5186 (2004).

d.25] A. Delia Villa, S. Enoch, G. Taycb, V. Pierro, V. Galdi, and F. Capolino,

Phys. Rev. Lett. 94, 183903 (2005).

d.26] B. Djafari-Rouhani, Y. Pennec, and J.O. Vasseur, private communication (2004).

d.27] R.D. Doolittle and H. Ueberall, J. Acoust. Soc. Am. 39, 272 (1966).

d.28] R.D. Doolittle and H. Ueberall, J. Acoust. Soc. Am. 43, 1 (1967).

d.29] J.M. Drake and A.Z. Genack, Phys. Rev. Lett. 63, 259 (1989).

d.30] M. Duneau and M. Audier, Acta Cryst. A55, 746 (1999).

d.31] E.N. Economou, Green's functions in quantum physics, Springer Verlag, Berlin

(1979).

d.32] B. Engquist and A. Majada, Math. Comp. 31, 629 (1977).

d.33] M.J. Escuti and G.P. Crawford, Opt. Eng. 43, 1973 (2004).

d.34] J.J. Faran, J. Acoust. Soc. Am. 23, 405 (1951).

d.35] Z. Feng, X. Zhang, Y.Q. Wang, Z.Y. Li, B. Cheng, and D.Z. Zhang,

Phys. Rev. Lett. 94, 247402 (2005).

104

[d.36] R.P. Feynman, R.B. Leighton, and M. Sands, The Feynman Lectures on Physics

Vol. II, Addison-Wesley Publishing Company, Reading Massachusetts (1964).

fd.37] F. Gähler and J. Rhyner, J. Phys. A: Math. Gen. 19, 267 (1986).

[d.38] R.C. Gauthier and K. Mnaymneh, Opt. Comm. 13, 1986 (2005).

|d.39] W. Gellermann, M. Kohomoto, B. Sutherland, and P. C. Taylor,

Phys. Rev. Lett. 72, 633 (1994).

[d.401 C. Gerthsen and H. Vogel, Physik. 17te Auflage, Springer, Berlin (1993).

fd.41] C. Godrèche and J.M. Luck, J. Phys. A: Math. Gen. 23, 3769 (1990).

[d.42] C. Goffaux, J. Saiichoz-Dchcsa, A.L. Yeyati, Ph. Lambin, A. Khelif, J.O. Vasseur,

and B. Djafari-Rhouhani, Phys. Rev. Lett. 88, 225502 (2002).

[d.43] S.P. Gorkhali, J. Qi, and G.P. Crawford, J. Opt. Soc. Am. B 23, 149 (2006).

[d.44] P. Gummelt, Geometr. Dedic. 62, 1 (1986).

[d.45l B.C. Gupta and Z. Ye, Phys. Rev. E 67, 036603 (2003).

[d.46l J. Hartwig, S. Agliozzo, J. Baruchel, R. Colella, M. deBoissieu, J. Gastaldi,

H. Klein, L. Mancini, and J. Wang, J Physics D 34, A103 (2001).

[d.47j M. Hase, H. Miyazaki, M. Egashira, N. Shinya, K.M. Kojima, and S.I. Uchida,

Phys. Rev. B 66, 214205 (2002).

[d.48] T. Hattori, N. Tsurumachi, S. Kawato, and H. Nakatsuka, Phys. Rev. B 50, 4220

(1994).

[d.49] S. He and J.D. Maynard, Phys. Rev. Lett. 62, 1888 (1989).

[d.50] V. Heine, Surf. Sei. 2, 1 (1964).

[d.51] M. Hoffe and M. Baake, Z. Kristallogr. 215, 441 (2000).

[d.52] N. Horiuehi, Y. Segawa, T. Nozokido, K. Mizuno, and H. Miyazaki, Opt. Lett. 29,

1084 (2004).

[d.53] R.M. Hornreich, M. Kugler, S. Shtrikman, and C. Sommers, J. Phys. 1 France 7,

509 (1997).

[d.54] C. Janot, Quasicrystals : a primer - Second ed., Clarendon Press, Oxford (1997).

fd.55] T. Janssen, Acta Crystallogr. A 42, 261 (1986).

105

[d.56] X.Y. Jiang, Y.G. Zhang, S.L. Feng, K.C. Huang, Y.H. Yi, and J.D. Joannopoulos,

Appl. Phys. Lett. 86, 1110 (2005).

[d.57] C.J. Jin, B.Y. Cheng, B.Y. Man, Z.L. Li, and D.Z. Zhang, Phys. Rev. B 61,

107C2 (2000).

[d.58] C.J. Jin, B.Y. Cheng, B.Y. Man, Z.L. Li, and D.Z. Zhang, Appl. Phys. Lett. 75,

1848 (1999).

[d.59] E. Jurdik et ed., Phys. Rev. B 69, 201102(R) (2004).

[d.60] M. Kafesaki and E.N. Economou, Phys. Rev. B 52, 13317 (1995).

[d.61] M.A. Kaliteevski, S. Brand, R.A. Abram, T.F. Krauss, R. DeLa Rue, and P. Mil¬

lar, Nanotechnology 11, 274 (2000).

[d.62] M.A. Kaliteevski, V.V. Nikolaev, R.A. Abram, and S. Brand, Opt. Speetr. 91,

109 (2001).

[d.63] M.A. Kaliteevski, D.M. Boggs, S. Brand, R.A. Abram, and V.V. Nikolaev,

Phys. Rev. E 73, 05G616 (2006).

fd.64] S. Kanoko, K. Edagawa, R. Tamura, and Y. G. So, Phil. Mag., submitted (2007).

[d.65] A. Khelif, B. Djafari-Rhouhani, J.O. Vasseur, P.A. Deymier, Ph. Lambin, and

L. Dobrzyns*i, Phys. Rev. B 65, 174308-1 (2002).

[d.66] A.D. Klironomos and E.N. Economou, Solid State Commun. 105, 327 (1997).

[d.67] M. Kohmoto and B. Sutherland, Phys. Rev. B 34, 3849 (1986).

fd.68l M. Kohmoto, B. Sutherland, and K. Iguchi, Phys. Rev. Lett. 58, 2436 (1987).

[d.69] M.S. Kushwaha, P. Halevi, L. Dobrzynski, and B. Djafari-Rouhani,

Phys. Rev. Lett. 71, 2022 (1993).

[d.70] M.S. Kushwaha, Appl. Phys. Lett. 70, 3218 (1997).

[d.71] M.S. Kushwaha and B. Djafari-Rouhani, J. of Sound and Vibration 218, 697

(1998).

[d.72] Y. Lai, X. Zhang, and Z.Q. Zhang, J. Appl. Phys. 91, 6191 (2002).

fd.73] Y. Lai, Z.Q. Zhang, C.H. Chan, and L. Tsang, Phys. Rev. B 74, 054305 (2006).

d.74] A. Lakhtakia, V.V. Varadan, and V.K. Varadan, J. Acoust. Soc. Am. 83, 1267

(1988).

106

fd.75] L.D. Landau and E.M. Lifshitz, Theory of elasticity - Third English edition,

Butterworth-Heinemann, Oxford (2002).

fd.76] A.V. Laviinenko, S.V. Zhukovsky, K.S. Sandomirski, and S.V. Gaponenko,

Phys. Rev. B 65, 036621 (2002).

fd.77] A. Ledeimann, L. Cademartiii, M. Hermatschweiler, C. Toninelli, G.A Ozin,

D.S. Wiersma, M. Wegener, and G. von Freymann, Nature Materials (London)

5, 942 (2006).

[d.78] D. Levine and P.J. Steinhardt, Phys. Rev. B 34, 596 (1986).

[d.79] M. Li, Y. Liu, and Z.Q. Zhang, Phys. Rev. B 61, 16193 (2000).

[d.80] Z. Liu, C.T. Chan, P. Sheng, A.L. Goertzen, and J.H. Page, Phys. Rev. B 62,

2446 (2000).

[d.81J Z. Liu, X. Zhang, Y. Mao, Y.Y. Zhu, Z. Yang, C.T. Chan, and P. Sheng, Science

289, 1734 (2000).

fd.82] J-M. Lourtioz, H. Benistiy, V. Berger, J-M. Gerard, D. Maystre, and A. Tchel-

nokov, Photonic Crystals Towards Nanoscale Photonic Devices. Springer Berlin

(2005).

[d.83] J.M. Luck, C. Godrèche, A. Janner, and T. Janssen, J. Phys. A 26, 1951 (1993).

[d.84] D. Lusk and R. Placido, Thin Solid Films 492, 226 (2005).

[d.85] B.Q. Ma, T. Wang, Y. Sheng, P.G. Ni, Y.Q. Wang, B.Y. Cheng, and D.Z. Zhang,

Appl. Phys. Lett. 87, 51103 (2005).

[d.86] W. Man, M. Megens, P.J Steinhardt, and P.M. Chaikin, Nature 436, 993 (2005).

[d.87] R. Meilin, K. Bajema, R. Claike, F.Y. Juang, and P.K. Bhattacharya,

Phys. Rev. Lett. 55, 1768 (1985).

[d.88] G. Mie, Ann. Phys. (Leipzig) 25, 377 (1908).

d.89] U. Mizntani, Introduction to the electron theory of metals, Cambridge University

Press, Cambridge (2001).

d.90] J.A. Monsoriu, C.J. Zapata-Rodrigue/, E. Silvestre, and W.D. Furlan,

Opt. Comm. 252, 46 (2005).

d.91] A. Montalbän, V.R. Vela&co, J. Tutor, and F.J. Fernandez-Vehcia, Phys. Rev. B

70, 132301 (2004).

107

[d.92] L. Moretti, I. Rea, L. Rotiroti, I. Rendina, G. Abbate, A. Marino, and L. Do

Stefano, Opt. Expr. 14, 6264 (2006).

[d.93] Laser Direct-Write Processing, MRS Bulletin 32, No. 1 (2007).

[d.94j G. Mur, IEEE Trans. Electromagn. Compat., 23, 377 (1987).

fd.95] G.G. Naumis and J.L. Aragon, Z. Kristallogr. 218, 397 (2003).

fd.96l M. Notomi, H. Suzuki, T. Tamamura, and K. Edagawa, Phys. Rev. Lett. 92,

123906 (2004).

fd.97] K. Nozaki, A. Nakagawa, D. Sano, and T. Baba, IEEE J. Selected Topics in

Quantum Eletronics 9, 1355 (2003).

[d.98] S. Ostlund and R. Pandit, Phys. Rev. B 29, 1394 (1984).

[d.99] J.H. Page, H.P. Schriemer, A.E. Bailey, and D.A. Weitz, Phys. Rev. E 52, 3106

(1995).

[d.100] J.H. Page, H.P. Schriemer, I.P. Jones, P. Sheng, and D.A. Weitz, Pysica A 241,

64 (1997).

[d.1011 J.H. Page, A.L. Goertzen, S. Yang, Z. Liu, C.T. Chan, and P. Sheng, NATO ASI

on Photonic Crystals and Light Localization, Crete (2000).

|d,102] G.J. Parker, M.D.B. Charlton, M.E. Zoorob, J.J. Baumberg, M.C. Netti, and

T. Lee, Phil. Trans. R. Soc. A 364, 189 (2006)

[d.103] R.W. Peng, X.Q. Huang, E. Qiu, M. Wang, A. Hu, S.S. Jiang, and M. Mazzer,

Appl. Phys. Lett. 80, 3063 (2002).

[d.l04j R.E. Prange, D.R. Grempel, and S. Fishman, Phys. Rev. B 28, 7370 (1983).

|d,105l I.E. Psarobas, N. Stefanou, and A. Modinos, Phys. Rev. B 62 (1), 278 (2000).

[d.1061 I.E. Psarobas, A. Modinos, R. Sainidou, and N. Stefanou, Phys. Rev. B 65 (1),

064307 (2001).

[d.1071 C. Radin, Annal. Math. 139, 661 (1994).

[d.108] P. Repetowicz and J. Wolny, J. Phys. A: Math. Gen. 31, 6873 (1998).

[d.109] P. Repetowicz, U. Grimm, and M. Schreiber, Phys. Revl. B 58, 13482 (1998).

[d.110] R. Riklund and M. Sevcrin, J. Phys. C 21, 3217 (1988).

[d.llll C. Rockstuhl, U. Peschel, and F. Lederer, Opt. Lett. 31, 1741 (2006).

108

[d.112] Y. Roichman and D.G. Grier, Opt. Expr. 13, 5434 (2005).

fd.113] J. Romero-Vivas, D.N. Chigrin, A.V. Lavrinenko, and C.M. Sotomayor-Torres,

Phys. Status Solidi (a) 202, 997 (2005).

[d.114] D.R. Roper, D.M. Beggs, M.A. Kaliteevski, S. Brand, and R.A. Abram, Journal

of Modern Optics 53, 407 (2006).

[d.115] R. Sainidou and N. Stefanou, Phys. Rev. B 66, 212301 (2002).

[d.H6j R. Sainidou and N. Stefanou, Phys. Rev. B 69, 064301 (2004).

[d. 117] R. Sainidou, N. Stefanou, I.E. Psarobas, and A. Modinos, Computer Physics

Communications 166, 197 (2005).

[d.118] R. Sainidou, N. Stefanou, and A. Modinos, Phys. Rev. Lett. 94, 205503 (2005).

[d.H9j R. Sainidou, N. Stefanou, I.E. Psarobas, and A. Modinos, Z. Kristallogr. 220,

848 (2005).

[d.120] .I.V. Sanchez, D. Caballero, R. Mârtinez-Sala, J. Sânchez-Dehesa, F. Meseguer,

J. Llinares, and F. Gâlvez, Phys. Rev. Lett. 80 (24), 5325 (1998).

[d.121] L. Sanchis, A. Hakansson, F. Cervera, and J. Sanchéz-Dehesa, Phys. Rev. B 67,

035422 (2003).

fd.1221 D. Shechtman, I. Blech, D. Gratias, and J.W. Cahn, Phys. Rev. Lett, 53, 1951

(1984).

[d.l23j P. Sheng, Introduction to Wave Scattering, Localization and Mesoscopic Phenomena,

Academic Press, San Diego (1995).

fd.124] P. Sheng, X.X. Zhang, Z. Liu, and C.T. Chan, Physica B 338, 201 (2003).

fd.125] P. Sheng and C.T. Chan, Z. Kristallogr. 220, 871 (2005).

[d.126] M. Shen and W. Cao, Appl. Phys. Lett. 75 (23), 3713 (1999).

[d,127j C. Sibilia, LS. Nefedov, M. Scalora, and M. Bertolotti, J. Opt. Soc. Am. B 15,

1947 (1998).

[d.128] M.M. Sigalas and E.N. Economou, J. Appl. Phys. 75 (6), 2845 (1994).

Id.l29j M.M. Sigalas, J. Appl. Phys. 84 (6), 3026 (1998).

[d.130] M. Sigalas, M.S. Kushwaha, E.N. Economou, M. Kafesaki, I.E. Psarobas, and

W. Steurer, Z. Kristallogr. 220, 765 (2005).

109

d.131] J.E.S. Socolar and P.J. Steinhardt, Phys. Rev. B 34, 617 (1986).

d.132] J.E.S. Socolar, Phys. Rev. B 39, 10519 (1989).

d.133] CM. Soukoulis, Photonic Band Gap Materials, Kluwer Academic Publishers, Dor¬

drecht (1996).

d.135] N. Stetiger, J-L. Rehspringer, and C. Hirlimann, J. Lumin. 21, 278 (2006).

d.136] W. Steurer and T. Haibach in: Z. M. Stadnik (Ed.), Physical properties of qua-

sicrystals, Springer, Berlin, Heidelberg (1999).

d.137] W. Steurer and T. Haibach, Acta. Cryst. A 55, 48 (1999).

d.138] W. Steurer, Mat. Sei. Eng. A 294, 268 (2000).

d.139] W. Steurer, Z. Kristallogr. 219, 391 (2004).

d.l40j W. Steurer, Z Kristallogr. 221, 402 (2006).

d.141] D. Sutter, G. Krauss, and W. Steurer, MRS Proceedings 805, 99 (2004).

d.1421 D. Sutter and W. Steurer, Phys. Status Solidi C 1, 2716 (2004).

d.1431 D. Sutter-Widmer, S. Deloudi, and W. Steurer, Phil. Mag. 87, 3095 (2007).

d.144] D. Sutter-Widmer and W. Steurer, Phys. Rev. B 75, 134303 (2007).

d.145] D. Sutter-Widmer, S. Deloudi and W. Steurer, Phys. R,ev. B 75, 094304 (2007).

d.146] A.P. Sutton, Electronic structure of materials, Oxford university press, Oxford

(1993).

d.147] A. Taflove and S.C. Hagness, Computational electrodynamics : the finite-difference

time-domain method - 2nd ed., Artech House, Boston (2000).

d.148] W.Y. Tarn, Appl. Phys. Lett. 89, 251111 (2006).

d.l49j S. Tamura and J.P. Wolfe, Phys. Rev. B 36, 3491 (1987).

d.150] H. Tsunetsugu, T. Fujiwara, K. Ueda, and T. Tokihiro, Phys. Rev. B 43, 8879

(1991).

d.151] J.O. Vasseur, B. Djafari-Rouhani, L. Dobrzynski, M.S. Kushwaha, and P. Halevi,

J. Phys. Condens. Matter 6, 8759 (1994).

d.152] J.O. Vasseur, P.A. Deymier, M. Chenni, M. Djafari-Rouhani, L. Dobrzynski, and

D. Prévost, Phys. Rev. Lett. 86, 3014 (2001).

110

[d.153] J.O. Vasseur, P.A. Dcymier, A. Khelif, Ph. Lambin, B. Djafari-Rouhaiii,

A. Akjouj, L. Dobrzynski, N. Fettouhi, and J. Zemmouri, Phys. Rev. E 65, 56608

(2002).

d.1541 M.S. Vasconcelos and E.L. Albuquerque, Phys. Rev. B 59, 11128 (1999).

d.l55j V.R. Velasco, R. Perez-Alvarez, and F. Garcia-Moliner, J. Phys.: Condens. Mat¬

ter 14, 5933 (2002).

d.l56j K. Wang, S. David, A. Chelnokov, and J.M. Louritioz, Journal of Modern Optics

50, 2095 (2003).

d.157] K. Wang, Phys. Rev. B 73, 235122 (2006).

d.158] X. Wang, C.Y. Ng, W.Y. Tarn, CT. Chan, and P. Sheng, Adv. Mater. 15, 1526

(2003).

d.159] Y. Wang, X. Hu, X. Xu, B. Cheng, and D. Zhang, Phys. Rev. B 68, 165106

(2003).

d.l60j Y. Wang, J. Liu, B. Zhang, S. Feng, and Z.Y. Li, Phys. Rev. B 73, 155107 (2006).

d.1611 Y. Weisskopf, S. Burkardt, M. Erbudak, and J.N. Longchamps, Surf. Sei. 601,

544 (2007).

d.162] F. Wu, Z. Liu, and Y. Liu, Phys. Rev. E 66, 046628 (2002).

d.163] E. Yablonovitch, Phys. Rev. Lett. 58, 2059 (1987).

d.164] A. Yamamoto, Acta Cryst. A 52, 509 (1996).

d.1651 Y. Yang, S.H. Zhang, and G.P. Wang, Appl. Phys. Lett. 88, 51104 (2004).

d.1661 K.S. Yee, IEEE Trans. Antennas Propagat. 14, 302 (1966).

d.167] J. Zarbakhsh, F. Hagmann, S.F. Mingaleev, K. Busch, and K. Hingerl,

Appl. Phys. Lett. 84, 4687 (2004).

d.168] X. Zhang, Z.Q. Zhang, and CT. Chan, Phys. Rev. B 63, 081105(R) (2001).

d.l69j Z.S. Zhang, B. Zhang, J. Xu, K. Xu, Z.J. Yang, Z.X. Qin, T.J. Yu, and D.P. Yu,

Appl. Phys. Lett. 88, 171103 (2006).

d.170] X. Zhang, Phys. Rev. B 75, 024209 (2007).

d.1711 S.N. Zhu, Y.Y Zhu, Y.Q. Qin, H.F. Wing, C.Z. Ge, and N.B. Ming,

Phys. Rev. Lett. 78, 2752 (1997).

11]

[d.172] M.E. Zoorob, M.D.B. Charlton, G.J. Parker, J.J. Baumberg, and M.C. Netti,

Nature 404, 740 (2000).

Acknowledgements

I am deeply indebted to Prof. W. Steurer for his support of this thesis, his constant

participation, his many ideas and suggestions as well as his patience. I much enjoyed the

freedom he granted me in my work.

Many members of the laboratory of crystallography earn my gratitude. They provided

first of all a friendly ambience at the laboratory but also constant support with whatever

problems there were.

We thank Dr. J.O. Vasseur and his collaborators at Université de Lille for their introduc¬

tory advises and provision of FDTD codes. We also thank him for acting as co-referee of

this thesis.

We thank Dr. U.G. Grimm for surveying this thesis.

We thank Dr. Y. Psarobas at University of Athens for fruitful discussions and collaboration

in the sphere project.

We thank Dr. R. Sainidou at University of Athens for discussions.

I thank the students which have contributed with semester works or thesis: Edith Fuchs-

berger, Stefan Waldburger, Angela Furier. Pedro Neves and Patrick Itten.

My parents have provided unconditional support during all of my education. They have

much earned my deepest gratitude (not only for this, of course).

My thanks to my loving wife Barbara.

Curriculum Vitae

Name

First name

Date of birth

Place of birth

Citizen

Marital status

Current adress

Sutter - Widmer

Daniel

May 12 1978

Lausen (BL)

of Bretzwil (BL)

married to Barbara Sutter - Widmer

Triemlistrasse 186, CH-8047 Zürich

2003 - 2007

2002

1998 - 2003

1994 - 1997

1985 - 1994

Ph. D. studies, Lab. of Crystallography, ETH Zürich

Material science studies, Australian National University (Canberra

ACT)

Material science studies, ETH Zürich

Gymnasium in Liestal (BL)

Elementary school in Holstein and Progymnasium in Oberdorf (BL)

2003 - 2007

2003

2002 - 2003

Ph. D. thesis "phononic quasicrystals" at the laboratory of crys¬

tallography ETHZ under supervision of Prof. W. Steurer.

Diploma thesis "Phononische Kristalle" at the laboratory of crys¬

tallography ETHZ under supervision of Prof. W. Steurer.

Semester thesis "Kristallstruktur von magnetoplastischen

NiMnGa-Legicrungen" at the institute of applied physics ETHZ

under supervision of Prof. G. Kostorz.

Publications

D. Sutter, G. Krauss, and W. Steurer, MRS Proceedings 805, 99 (2004).

D. Sutter and W. Steurer, Phys. Status Solidi C 1, 2716 (2004).

D. Sutter-Widmer, S. Deloudi, and W. Steurer, Phil. Mag. 87, 3095 (2007).

D. Sutter-Widmer and W. Steurer, Phys. Rev. B 75, 134303 (2007).

D. Sut ter-Widmer, S. Dcloudi and W. Steurer, Phys. Rev. B 75, 094304 (2007).

W. Steurer and D. Sutter-Widmer, Sol. State Phen. 130, 33 (2007).

W. Steurer and D. Suttor-Widmor, J. Phys. D: Appl. Phys. 40, R229 (2007).