chemical society reviews - stuba.skszolcsanyi/education/files/the nobel...the determination of their...

12
ISSN 0306-0012 0306-0012(2012)41:20;1-A www.rsc.org/chemsocrev Volume 41 | Number 20 | 21 October 2012 | Pages 6709–6848 Themed issue: Quasicrystals Chemical Society Reviews Guest editor: Walter Steurer TUTORIAL REVIEW Walter Steurer Why are quasicrystals quasiperiodic? Downloaded on 27 September 2012 Published on 22 May 2012 on http://pubs.rsc.org | doi:10.1039/C2CS35063G View Online / Journal Homepage / Table of Contents for this issue

Upload: others

Post on 20-Jul-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

ISSN 0306-0012

0306-0012(2012)41:20;1-A

www.rsc.org/chemsocrev Volume 41 | Number 20 | 21 October 2012 | Pages 6709–6848

Themed issue: Quasicrystals

Chemical Society Reviews

Guest editor: Walter Steurer

TUTORIAL REVIEWWalter SteurerWhy are quasicrystals quasiperiodic?

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

GView Online / Journal Homepage / Table of Contents for this issue

Page 2: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

This journal is c The Royal Society of Chemistry 2012 Chem. Soc. Rev., 2012, 41, 6719–6729 6719

Cite this: Chem. Soc. Rev., 2012, 41, 6719–6729

Why are quasicrystals quasiperiodic?w

Walter Steurer*

Received 5th March 2012

DOI: 10.1039/c2cs35063g

In the past two decades significant progress has been made in the search for stable quasicrystals,

the determination of their structures and the understanding of their physical properties.

Now, quasiperiodic ordering states are not only known for intermetallic compounds, but also for

mesoscopic systems such as ABC-star terpolymers, liquid crystals or different kinds of colloids.

However, in spite of all these achievements fundamental questions concerning quasicrystal formation,

growth and stability are still not fully answered. This tutorial review introduces the research in these

fields and addresses some of the open questions concerning the origin of quasiperiodicity.

1 Introduction

The first intermetallic compound identified as quasicrystal

(QC)z was a rapidly solidified metastable Al–Mn phase with

icosahedral diffraction symmetry.2 Dan Shechtman obtained it

on 8 April 1982 employing the melt-spinning method, which

allows the quenching of melts within milliseconds.

A few years later, it was shown that less rapidly solidified

Al–Mn forms a metastable decagonal QC (DQC). Only in the

case of even slower cooling rates a stable phase forms, which is

periodic then.3 The faster crystallization of the icosahedral QC

(IQC), compared to the periodic phase, was explained by its

lower nucleation barrier because in the melt it is already existing in

the icosahedral short-range order.4–6 However, the easy nucleation

of clusters with icosahedral symmetry does not necessarily imply

quasiperiodic growth. The existence of QC approximants

clearly demonstrates that the same kind of structural subunits

(in the following called ‘cluster’7,8 or ‘quasi-unit cell’9) can

basically constitute both quasiperiodic and periodic structures.

In the year 1985, the first stable QC was identified in the system

Al–Cu–Li,10 which has been overlooked in previous studies. This

discovery showed that quasiperiodic order can be a thermodynamic

equilibrium state of matter and motivated a systematic search for

other stable QC in intermetallic systems, guided by certain expected

valence-electron/atom ratios and the existence of approximants,

if any.11 The strongly differing compositional stability fields of

the mostly ternary QC found so far are shown in Fig. 1.

The temperature stability ranges are differing largely as well,

from a few to several hundred degrees. Unfortunately, due to

the sluggish kinetics of phase transformations of intermetallics

at low temperatures, it cannot be decided whether or not the

quasicrystalline ordering state can be a ground state of matter.

Quasiperiodic structures and approximants have also been

observed in some less metallic materials such as rapidly

solidified borides13 or conventionally synthesized tellurides.14

However, in both cases particular preferred structural subunits

favor random-tiling-like arrangements with non-crystallographic

10- and 12-fold symmetry, respectively.

In recent years, also an increasing number of ‘soft quasicrystals’

have been identified with self-assembled quasiperiodic structures

on the mesoscale. Examples are supramolecular dendrimer

liquid QC (LQC),15 ABC-star terpolymers16 and surfactant-

coated metallic nanoparticles in solution.17 Furthermore, forced

quasiperiodic structures have been found for polystyrene particles

Laboratory of Crystallography, ETH Zurich, Wolfgang-Pauli-Strasse 10,CH-8093 Zurich, Switzerland. E-mail: [email protected];Fax: +41 44 632 1133; Tel: +41 44 632 6650w Part of a themed issue on Quasicrystals in honour of the 2011 NobelPrize in Chemistry winner, Professor Dan Shechtman.

Walter Steurer

Walter Steurer studied chemistryand received his PhD at theUniversity of Vienna. In 1980,he moved to the University ofMunich, Germany, where heconcluded his habilitationthesis in 1987. From 1992until 1993, he was a professorof crystallography at theUniversity of Hanover,Germany. After declining callsto the Universities of Hamburgand Munich, Germany, he hasbeen a professor of crystallo-graphy at the ETH andUniversity of Zurich since

1993. At present, his research interests comprise structuralstudies of quasicrystals and their phase transformations, themodeling of order/disorder phenomena, and higher-dimensionalcrystallography.

z Originally it was called ‘icosahedral phase’; the term ‘quasicrystal’ wascoined in 1984 by Levine and Steinhardt:1 ‘‘A quasicrystal is the naturalextension of the notion of a crystal to structures with quasiperiodic,rather than periodic, translational order’’. As described in Section 2.2,we focus on QC with non-crystallographic symmetry.

Chem Soc Rev Dynamic Article Links

www.rsc.org/csr TUTORIAL REVIEW

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 3: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

6720 Chem. Soc. Rev., 2012, 41, 6719–6729 This journal is c The Royal Society of Chemistry 2012

in solution under the action of laser interference fields18 as well

as for colloidal micelles in a shear rheometer19 (for recent

reviews see the studies of Dotera20 and Ungar et al.21).

In spite of three decades of research into QC resulting in

more than 10 000 publications so far, there are still funda-

mental questions open. Some of them will be addressed in the

following from a structural point of view:

�Which factors determine whether a structure grows periodi-

cally or quasiperiodically?

� What guides the growth of a quasiperiodic structure?y�What stabilizes a quasiperiodic structure over energetically

competing periodic ones?

This tutorial review is not intended to be exhaustive and to

cover all important contributions to this field. It is just

intended to be an introduction into some of the most impor-

tant, and sometimes controversially discussed, open problems

of QC research. Of course, the selection and presentation of

topics is biased by the interests of the author. For more in-depth

information see the recently published books by Steurer and

Deloudi,12 Janssen et al.23 and Mizutani.24

2 Generalized crystallography

2.1 Structure determining factors

Crystal structures can be seen resulting as a compromise

between short-range order (sro) and long-range order (lro).

Every single atom tries to take on its energetically most

favorable atomic environment type (AET) at the same time,

which not only constrains the formation of the preferred AET

of atom X with that of the neighboring atom Y, which is

already part of the AET of X, but also all these individual

AET can be geometrically very unfavorable for packing them

space-filling in a crystal structure.

The simplest cases, where every atom has exactly the same

AET, are the cubic and hexagonal close-packed structures. They

have the highest possible coordination for equal spheres as well as

the maximum space filling, locally and globally. However, much

less metallic elements than one may expect crystallize in such

structures, in particular, at higher pressures. The reason is

anisotropic interaction potentials caused, for instance, by covalent

bonding contributions, particular electronic and/or magnetic

interactions and/or specific phononic contributions to entropy.

Crystal-chemical (atomic interactions) and geometrical

constraints (stoichiometry and atomic size ratios) for structure

formation can become quite complex in the case of multinary

intermetallics. This limits the number of different compounds

that are possible in a given intermetallic system. Indeed, this

number decreases drastically going from binary to ternary or

even higher multinary systems. The more complex the stoichio-

metry of an intermetallic compound becomes, the larger the

number of different AET becomes and the more difficult

their packing. Metastable metallic glasses (amorphous metals)

represent the cases where all the irregular AET already present

in the melt of a multinary alloyz cannot be properly modified

on the timescale of solidification for achieving lro. Thermal

equilibration then usually leads to phase separation, where the

individual phases have rather simple structures.

Complex structures of intermetallic phases frequently show

larger structural subunits (‘clusters’), each of them constituted

by several AET. The cluster formation can drastically reduce

the number of different structural units to be optimized for

packing (see, e.g., Dshemuchadse et al.25). Clusters are usually

of higher symmetry than their constituting AET and their

chemical composition is also closer to that of the compound.

For an example see Fig. 2, where the cluster structure of the

cubic 2/1-approximant Cd76Cd1326 of the Tsai-type IQC is shown

together with the cluster-constituting CaCd16 polyhedron.

In the case of entropy-driven substitutional (chemical)

and/or displacive disorder, individual AET may become strongly

distorted. In contrast, structural disorder can be easily accom-

modated in inner cluster shells without influencing the way of

cluster packing. Another kind of disorder is caused by phason

fluctuations, i.e., low-energy excitations in QC that lead to short-

distance atomic jumps in double-well potentials (see Fig. 3(c)).

With a few of them correlated, virtual jumps of whole clusters

can result (see Fig. 9 of Coddens and Steurer27). To an extent,

this kind of disorder (‘phasonic disorder’) can result in random-

tiling-like structures, where energy differences between allowed

and forbidden tile configurations (in ideal quasiperiodic tilings)

are overcompensated by the entropy gain.

Fig. 1 Compositional stability fields of (top) decagonal and (bottom)

icosahedral QC (adapted from ref. 12). Note the different concen-

tration ranges of the diagrams.

y Roger Penrose, the creator of the quasiperiodic Penrose tiling,reminisces: ‘‘. . .in the late 1970s and early 80’s I had often been askedto give lectures on these tiling patterns, and a question frequentlyposed to me after the lecture might be: ‘Does this not mean that thereis a whole new area of crystallography opening up, with pentagonaland icosahedral symmetry allowed?’ My normal response would be:‘In principle yes; but how on earth would Nature do it?’. . .thespontaneous growth of large regions of such quasi-crystalline five-foldsymmetric substances had seemed to me virtually insurmountable.’’22

z There is a wealth of experimental evidence about the preformationof polytetrahedral clusters, such as pentagonal bipyramids (decahedra)or icosahedra, in supercooled metallic melts at compositions whereQC form after solidification; see, for instance, the system Al–Cu–Coreported by Roik et al.6 and citations therein.

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 4: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

This journal is c The Royal Society of Chemistry 2012 Chem. Soc. Rev., 2012, 41, 6719–6729 6721

The QC-constituting clusters usually show ‘non-crystallo-

graphic’ 5-, 8-, 10- or 12-fold rotational symmetry. The densest

possible packings of such clusters can be derived employing

the higher-dimensional approach (see Section 2.2). Here, the

basic underlying idea is illustrated on the example of a simple

1D quasiperiodic structure, the Fibonacci sequence (FS), and

its approximants (Fig. 3). The vertices of the FS are separated

by distances L and S, with L = tS. If the composition of the

FS or an approximant differs from that of its constituting

cluster, then it is obvious that it can be only achieved by

allowing S-type overlaps. For the derivation of the sequence of

overlaps that leads to the smallest deviation of the local from

the global stoichiometry, we use the fact that the cluster (LS)

can be obtained by the projection of the edges of a square in

proper orientation. As shown in Fig. 3(b), the most homo-

geneous packing of (LS) clusters building the FS with overall

composition LtS then corresponds to the packing of 2D unit

cells in a strip cut out of a 2D lattice.

Contrary to the structures of hard QC, those of soft QC are

not cluster-based and follow other self-assembly mechanisms

(see Section 4). Soft QC usually consist of just one type of

building element such as micelles or other colloidal particles.

In order to get something more complex than dense sphere

packings, these ‘mesoscopic atoms’ must have complex inter-

action potentials providing two length scales, for instance. In

the case of ABC-star terpolymers, additionally the stoichio-

metry AxByCz plays a role for the kind of structure formed by

the minimization of the interfacial area between the immiscible

polymers A, B and C.

Due to the specific interaction potentials on the mesoscopic

scale, the resulting self-assembled quasiperiodic structures all

show dodecagonal diffraction symmetry in contrast to the

5-fold symmetry prevalent in hard QC. The tetrahedrally close

packed (tcp) structures resemble decorated random square–

triangle tilings. Amazingly, simulations of dense packings of

hard tetrahedra29 and trigonal bipyramids30 show dodecagonal

diffraction symmetry and random-tiling-like arrangements of

the polyhedra centers as well.

2.2 Structural description of quasicrystals and their rational

approximants

Quasiperiodic structures are defined as having Fourier modules

of rank n> d, with d being the dimension of the structure. This

definition includes incommensurately modulated structures

(IMS) and composite structures (CS) with crystallographic

symmetries, as well. Since IMS and CS can be described

referring to one or more dD periodic basic structures, they

are fundamentally different from quasiperiodic structures with

non-crystallographic symmetry. Consequently, we mean by QC

only those quasiperiodic structures that have non-crystallographic

symmetry.

The classification of QC is based on their diffraction symmetry

(N-fold axial symmetry or icosahedral symmetry), on one hand,

and the type of clusters they consist of, on the other hand. In the

case of IQC, the three different structure types are based on

Mackay-, Bergmann- and Tsai-clusters, respectively. In the case

of DQC, we distinguish between structures with 2-, 4-, 6- or

8-layer periodicity, and whether they are Al-based or of

Fig. 2 Shells of the Tsai-cluster in the cubic 2/1-approximant

Cd76Ca13 (a = 25.339 A, Pa%3).26 (a) Cd20 dodecahedra each enclosing

one orientationally disordered Cd4 tetrahedron (not shown); (b) Ca12icosahedra around the Cd20 dodecahedra; (c) distorted Cd30 icosido-

decahedra connected via trigonal antiprisms to each other; (d) Cd80edge-decorated defective triacontahedra sharing oblate rhombohedra

with each other; (e) the cluster-constituting CaCd16 polyhedron with

the shells indicated, which are formed by assembling twelve of them.

Fig. 3 (a) 1D structures with overall composition LS (top), L2S

(middle) and LtS, i.e., two approximants and the FS, respectively

(bottom); the decomposition into covering clusters of the type (LS) is

shown above each structure. (b) FS in the 2D description employing

the strip-projection method. The projection of the sequence of vertex

and edge connected shaded squares shown gives the FS plus some

additional vertices (gray) at (phason) flip positions in a double-well

potential (c). The width W of the strip defines the minimum distance

between projected lattice points as well as the unit tile sequence; in our

case, W is chosen so that the lowermost (gray) vertex of the yellow

square is outside the strip. (d) A perp-space shear of the lattice

(indicated by the arrow at right) leads to the (LSL) approximant

(modified from Steurer and Deloudi28).

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 5: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

6722 Chem. Soc. Rev., 2012, 41, 6719–6729 This journal is c The Royal Society of Chemistry 2012

the Zn–Mg–RE (RE—rare earth element) class.31 The latter

are related to the Frank–Kasper phases.

Basically, there are two ways of describing the structure of

QC and their rational approximants (for a short introduction

into the crystallography of QC see the review of Steurer and

Deloudi,32 for instance, and for a comprehensive textbook see

Steurer and Deloudi12). The tiling- or ‘quasilattice’-approach,

on one hand, is similar to the description of periodic structures

by lattices decorated with atoms. In contrast to lattices,

quasilattices have at least two different ‘unit cells’ (prototiles)

that can be decorated by atoms or clusters (Fig. 4).

Partially overlapping ‘covering clusters’ allow us to describe

the full quasiperiodic structure even by one type of cluster33

(‘quasi-unit cell’9). In the case of the FS, the covering cluster

corresponds to the sequence (LS) (Fig. 3), in the case of

the Penrose tiling, the Gummelt decagon plays this role

(Fig. 4 and 5).

The higher-dimensional or nD approach, on the other hand,

allows us to describe dD quasiperiodic structures based on

periodic nD hypercrystal structures (n> d). The nD embedding

space, V = VJ"V>, consists of the dD physical or par(allel)

spaceVJ and the (n� d)D perp(endicular) or complementary space

V>; n corresponds to the rank of the Fourier module of the

QC, i.e., the number of reciprocal basis vectors needed to

index the reflections of a diffraction pattern with integers only.

There are two equivalent methods, the strip-projection method

and the embedding method. In the strip-projection method,

illustrated in Fig. 3, those nodes of an nD lattice are selected

by a strip that give a quasiperiodic tiling, when projected

onto the par-space. The width W of the strip results from the

projection of one nD unit cell onto then perp-space.

The embedding method is based on the description of a dD

quasiperiodic structure as the proper irrational section of a

periodic nD hypercrystal structure (n > d). In reciprocal

space, this corresponds to a projection of the nD reciprocal

lattice onto the physical space.

Let us assume, now, that we have a 2D decagonal cluster,

which can be described by a Gummelt decagon with its specific

overlap regions. In a 2D packing, the decagon centers form a

subset T of a Z-module of rank n= 5. Since a Z-module of rank

5 can be created by projecting along a 5-fold axis a 5D hypercubic

lattice onto 2D par-space, the subset T corresponds to a

connected portion of the 5D hypercubic lattice. These relationships

are schematically illustrated in Fig. 5. For more details about

the nD cluster approach see the study of Steurer and Deloudi.28

2.3 Complexity and periodicity versus quasiperiodicity

All experimentally observed intermetallic phases in thermal

equilibrium are crystalline, with either periodic or aperiodic

crystal structures. They all are characterized by Fourier spectra

containing a dominant pure point component (set of Bragg

peaks) beside a, usually much weaker, continuous part such as

thermal, phason and disorder diffuse scattering.

Fig. 4 Electron density map of decagonal Al–Co–Ni projected down

the tenfold axis. Typical covering clusters (yellow decagons with blue

markings and white Penrose rhomb decoration), so-called ‘Gummelt

decagons’,33 are shown with their blue markings indicating the overlap

regions. The Gummelt decagons can be seen as projections of 5D

hypercubic unit cells as visualized in Fig. 5. In addition to the

Gummelt decagons, the two sets of inequivalent quasilattice lines are

shown (red) as well as a patch of a rhomb Penrose tiling (green)

(modified from Steurer and Deloudi28).

Fig. 5 (top) Two, 3D-face sharing, properly embedded 5D hypercubes

in combined par-/perp-space projections giving elongated rhombic

icosahedra and overlapping (area black outlined) Gummelt decagons

A and B, respectively. (bottom) The Gummelt decagons A, B and one

additional one, C, with two different types of permitted overlaps.

Green dots mark the shared vertices of the hypercubes related to the

decagons A and B, black circles those of B and C (modified from

Steurer and Deloudi28).

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 6: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

This journal is c The Royal Society of Chemistry 2012 Chem. Soc. Rev., 2012, 41, 6719–6729 6723

Before entering the discussion on why particular materials

are periodic and others are quasiperiodic, we first try to shortly

summarize the benefits of periodicity, which must be enormous

because almost all known crystals are periodic. Of course, real

structures are never strictly (quasi)periodic, but only on average.

In thermal equilibrium, the (quasi)periodicity is broken at least

by the finite size of the crystals, thermal vacancies, phonons,

and, particularly in complex intermetallics, frequently by

structural disorder. However, these deviations from periodicity

are usually small compared to interatomic distances.

Periodicity means that the energetically most favorable

structural subunit (cluster or unit cell) acts as a building block

leading to an energy- and entropy-optimized arrangement of

atoms. Moreover, periodicity usually results in, due to the

existence of Brillouin zones, simpler electronic band structures

with well-defined band gaps.

One measure of the complexity of a structure is its generalized

R-atlas, which describes the environment of a given atom within

a sphere of radius R. If each atom in a structure has exactly

the same R-atlas up to infinity, i.e., the number N of different

R-atlases equals one for any R, then the structure must be

periodic and consist of one kind of atoms, only. In the case of a

more complex, but still periodic structure, each of theM atoms

in the unit cell can have a different R-atlas. For the whole

infinite structure the number of different R-atlases is still finite,

N = M for any R. In the case of quasiperiodic structures,

however, the number of different R-atlases grows from R to

infinity, even for equally decorated vertices. From this point

of view, QC seem to be much more diverse and complex

than periodic crystals. According to the parsimony principle,

periodic structures should be generally more favorable than

quasiperiodic ones, consequently. This is the case, indeed.

However, the real complexity of QC is much lower than

indicated by their dD R-atlases. It is much more realistically

reflected in the nD description, where the number of different

nD R-atlases gets finite again for any R, reflecting the intrinsic

correlations present in quasiperiodic structures. For instance,

the 6D primitive hypercubic unit cell of the 3DAmmann8 tiling in

the higher-dimensional description contains just one ‘hyperatom’

resulting in the simplest possible R-atlas withN= 1 for any R.

Of course, it is difficult to quantitatively compare 3D and nD

R-atlases. The qualitative comparison, however, clearly shows

that the number of different R-atlases is finite for any R, in both

cases. Moreover, the nD R-atlases of QC seem to be much

simpler than many of the 3D ones of complex intermetallics.

Although for amorphous structures the R-atlases grow from

R to infinity, as well, it should be emphasized that QC are not

intermediate between amorphous alloys and crystals, they are

crystals! Even in the case of thermally equilibrated 3D QC that

can be described based on random tilings, the autocorrelation

function keeps a discrete part while it is continuous in the case

of amorphous materials.

Another measure of complexity for comparing periodic

structures is the number of atoms per unit cell, which for

almost all structures is much smaller than 400 (Fig. 6). QC

could be compared with each other by comparing the com-

plexity of their clusters or quasi-unit cells. However, all QC

known so far are of comparable complexity.

The number of intermetallic phases steeply decreases with

increasing number of atoms. The small peak at E450 atoms

per unit cell mainly stems from a set of structures based on

close packing of large clusters.25 There are only a few more

structures left with more than 1000 atoms per unit cell. The

so far largest structure of an intermetallic phase contains

23 134 atoms in its cubic unit cell with lattice parameter

a = 71.490 A.35

The steeply decaying frequency distribution can be inter-

preted by decreasing benefits of periodicity with increasing

unit cell dimensions, while the costs of maintaining periodicity,

by packing more and more different AET, increase. Benefits

of small unit cell periodicity are simple construction plans

based on the packing of only a small number of different

AET, a diffraction pattern with a high fraction of strong

reflections indicating the existence of well-defined atomic

layers parallel to lattice planes, repeating themselves with

small distances.

The limits of the complexity of intermetallic crystal structures are

also indicated by the number of binary, ternary and quaternary

intermetallic compounds known so far. From a combinatorial

point of view, the number of different crystal structures should

increase going from binary to higher multinary intermetallic

systems. The opposite is true as already mentioned in Section 2.1.

There are only a few truly quaternary phases known, which

are not just superstructures of ternary or binary phases such as

some Heusler phases or metastable ternary phases that are

stabilized by the addition of a fourth element. Similar to that

in the case of intermetallics with odd chemical compositions

Fig. 6 The number of intermetallic phases (on logarithmic scale) as a

function of the number of atoms per unit cell (bin size 10). The data of

the E43 000 compounds have been extracted from Pearson’s Crystal

Data34 (courtesy of J. Dshemuchadse). Approximately twenty compounds

with more than 1000 atoms have been omitted from the histogram to

save space. 8 Also known as 3D Penrose tiling.

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 7: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

6724 Chem. Soc. Rev., 2012, 41, 6719–6729 This journal is c The Royal Society of Chemistry 2012

(complex stoichiometry, strongly differing atomic radii), in

multinary phases many different AET are necessary to reach,

on average, the global chemical composition; and the more

different AET exist the more complex become their packings

and the larger the resulting unit cells.

Periodicity has to be strongly disfavored in order to get a

structure that is quasiperiodic or amorphous. How can this be

achieved? One way is by mixing several elements with strongly

differing atomic radii and rather rapid solidification of the

melt in order to avoid phase separation and crystallization;

thereby, amorphous materials can be obtained. Another way

is to choose elements forming endohedral clusters with non-

crystallographic symmetry whose densest packing is quasi-

periodic (see Fig. 5). These clusters must be energetically

favorable and they must allow specific cluster overlaps.

Furthermore, electronic stabilization of the structure seems

to be beneficial if not necessary.

An important feature, and used for the definition of crystals,

is that both dD periodic and quasiperiodic structures show

pure point (Bragg) spectra, corresponding to Fourier modules

of rank n. In the case of a periodic structure the rank equals

the dimension, n= d, in the case of a quasiperiodic structure it

is larger, n> d. The Bragg reflection density is finite in the first

case and infinite in the second. Related to the Bragg reflections

in reciprocal space, in quasiperiodic structures a framework of

‘quasilattice planes’ exists, which can be seen as interfaces

between adjacent parts of the structure (see Fig. 4 and 7). In

the case of quasiperiodic structures, the non-crystallographic

symmetry of the fundamental clusters coincides with the

framework of quasilattice planes (see Fig. 7). In the case of

its periodic approximant this is not the case.

The atomic layers on the quasilattice planes are important

for the existence and propagation of phonons and of electron

(Bloch) waves. Flat atomic layers correspond to strong Bragg

peaks and therewith to well defined Brillouin zones; puckering

decreases the intensity of Bragg peaks. Atomic layers also make

possible facetted 3D crystal growth: all facets that grow at the

same time independently from each other always fit together

where they meet forming the crystal edges. The framework of

atomic layers connects sro with lro, and the internal symmetry

of clusters with the global symmetry of the infinite structure.

In random-tiling based soft QC, quasilattice planes do not

exist; however, this does not play a role for the entropy-driven

ordering of LQC.

It is quite easy to transform a QC with non-crystallographic

symmetry into a structurally closely related aperiodic structure

with crystallographic symmetry. In reciprocal space, the

respective distortion of the quasiperiodic structure corre-

sponds to a change in the angles between the basis vectors.

This kind of transformation does neither change the rank of

the Fourier module nor does it make the structure periodic.

A rational approximant can be obtained within the nD approach

by a particular shear transformation of the hypercrystal structure

(see Fig. 3). Thereby, the rank n of the Fourier module is lowered

to n = d. In direct space, this structural phase transformation

could be achieved locally by a correlated rearrangement of atoms

(for a review see Steurer36) leading to a kind of nanodomain

structure. It needs a first-order transformation to transform a

QC into a periodic structure.

2.4 Model calculations

Model calculations are absolutely essential for gaining insight

into the factors controlling structure, formation and stability

of QC. Most commonly used are classical Monte Carlo (MC)

or Molecular Dynamics (MD) simulations employing specific

pair potentials for the interaction of atoms or particles (in the

case of mesoscopic QC). On a larger hierarchy level, calcula-

tions based on tile Hamiltonians37 have been proven useful for

the study of medium range order (mro). Quantum mechanical

calculations are computationally more expensive and therefore

even more restricted by the model system size, let alone the

periodicity conditions always needed. In other words, first

principles calculations suffer most from the periodicity condi-

tions because of the small possible model system size of a few

hundred atoms per unit cell. Taking into account typical

cluster sizes of E150 atoms, this corresponds to just a few

clusters per unit cell. Consequently, the quasiperiodic lro,

i.e., exactly the property that distinguishes QC from periodic

crystals, cannot be modeled properly. With other words, the

information obtained from all these calculations refers to

approximant model structures and not to quasiperiodic ones.

In classical MC or MD simulations a certain chemical

composition and density are predefined. This implicitly forces

a particular arrangement of the energetically favorable AET

into larger subunits (clusters) as well as a particular ordering

of the resulting clusters themselves. Whether periodic or

quasiperiodic order results in a lower free energy for a given

composition and density depends on the cluster packing

optimization rather than directly on pair potentials.

Fig. 7 Projections of spherical sections (100 A diameter) of the

structure of ico-Cd–Yb (data courtesy of H. Takakura) along (a) a

5-fold axis and (c) a 2-fold axis. In (b) and (d) the corresponding

projections of the periodic 1/1-approximant are depicted, i.e., along

the pseudo 5-fold and the 2-fold directions, respectively. In (c) and (d)

a few atomic layers are marked. In (c) their distances follow a FS, in

(d) their period equals half the lattice parameter.

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 8: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

This journal is c The Royal Society of Chemistry 2012 Chem. Soc. Rev., 2012, 41, 6719–6729 6725

This said, what kind of useful information can be obtained

employing the methods mentioned above? First principles

calculations can give insight into the local chemical bonding

of energetically favorable cluster shells. They can clarify

whether structure stabilizing pseudo-gaps in the electronic

density of states (eDOS) at the Fermi level originate from

hybridization of atoms in the innermost cluster shells or rather

from a Hume-Rothery mechanism (Fermi surface nesting).

However, it should be kept in mind that this information refers

rather to a low approximant than to a quasiperiodic system;

furthermore, disorder of any kind is not included in stability

calculations of this kind.

Although the model size is rather small, local disorders such

as phason flips can be studied quite well in classical MC and

MD simulations, which can be based on system sizes up to

millions of atoms. The strength of these methods is their

capability in investigating the role of pair potentials for the

stabilization of clusters and their preferred arrangements,

i.e., the local evolution of quasiperiodic order and the kind

and role of disorder even as a function of temperature.

In the case of mesoscopic systems MC and MD simulations

can give much deeper insight into the self-assembly mecha-

nisms of mesoscopic systems than in the case of intermetallics.

There, the interaction potential is given by the structure of the

micelles and does only change with density and temperature.

For instance, it was demonstrated by T = 0 MD simulations

that even in monatomic systems structures with decagonal and

dodecagonal diffraction symmetry can be stabilized if parti-

cular double-well potentials are used.38 One has to keep in

mind, however, that these were 2D model systems, only.

However, it is important to emphasize that the models that

work quite well for mesoscopic systems cannot be directly

transferred to intermetallic QC; the underlying mechanisms

are too different, in particular the kind of bonding and

electronic stabilization. This is reflected in the prevalence of

5-fold symmetry and good quasiperiodic order in the case of

hard QC and 12-fold symmetry and random-tiling-like order

in the case of soft QC. The pair potential for interactions

between micelles is exclusively defined by the individual

micelle and fully transferable between different structures. In

contrast, due to multibody interactions and the existence of

itinerant electrons, the pair potentials of atoms in intermetallic

phases strongly depend on the kind and arrangement of

neighboring atoms as well as on the structure as a whole.39

3 Intermetallic (‘hard’) quasicrystals

Up to date, stable intermetallic QC could be synthesized as

(pseudo)binary or ternary phases with either decagonal or

icosahedral diffraction symmetry, only. Furthermore, there is

one, probably stable, semimetallic dodecagonal QC known,

Ta1.6Te, with a lubricant-like layer structure,40,41 which can be

seen to take an intermediate position between hard and soft

QC. Furthermore, a number of metastable octagonal and

dodecagonal intermetallic phases have been prepared by rapid

solidification techniques.

From their electronic structure, two classes of hard QC can

be distinguished, Al–TM (TM—transition metal) based QC

where d-orbitals play a crucial role and Mg–Zn based QC

(Frank–Kasper type) with s- and p-electrons dominating

metallic bonding. While in the first case pair potentials with

Friedel oscillations seem to be appropriate for modeling,42 in

the latter case even simple Lennard-Jones potentials have been

found to stabilize QC structures (see, e.g., Roth and Henley43

and citations therein).

The question is why no hard QC with other non-crystallo-

graphic symmetries are known, in contrast to the findings for

soft QC (see Section 4). For instance, there are ternary borides

that can be seen as approximants of heptagonal QC,44,45

however, neither stable nor metastable hard QC with this

symmetry have been observed ever.

Concerning the stability of intermetallic QC, it is interesting

to note that only a few of them seem to be truly ternary phases.

Most of them are pseudobinaries, needing a particular amount

of a third element just for tuning the electron concentration,

stabilizing particular AET or increasing the entropy by

chemical (substitutional) disorder. For instance, the meta-

stable IQC Al–Mn can be stabilized by substituting a part of

Al and Mn by Pd. It is also remarkable that all quasiperiodic

structures determined so far show a clear cluster substructure,

with just a few topologically differing clusters. In the case of

decagonal and icosahedral phases the underlying quasilattice

(quasiperiodic tiling) is closely related to the 2D Penrose tiling

and the 3D Ammann tiling.

The typical structural units of decagonal quasicrystals can

be described based on the so-called Gummelt clusters33,46 that

are decorated in a particular way. In the case of icosahedral

quasicrystals, the structures can be seen to consist of subunits

which can be inscribed into triacontahedra. Both types of

clusters can be described as proper projections of nD hyper-

cubes, with n = 5 and n = 6, respectively.28

3.1 Decagonal quasicrystals (DQC)

So far, stable DQC have been found in ternary intermetallic

systems only, while metastable DQC have also been obtained

by rapid solidification of binary phases (Fig. 1) (for a review

on DQC see Steurer31). There are a few systems known such as

Al–Mn–Pd and Zn–Mg–Dy, where both stable DQC and IQC

with slightly different compositions exist.

The structures of DQC show 2D quasiperiodicity perpendicular

to the tenfold axis and periodicity along it. Geometrically, they

can be described as periodic stackings of quasiperiodic layers.

However, from a crystal-chemical point of view DQC are

by no means layer structures as indicated already by their

needle-like growth morphology. The interatomic distances

within layers and between layers are on the same scale.

Therefore, DQC should better be described as packings of

each other partially overlapping columnar clusters. Amazingly,

all structures of DQC analyzed so far can be described as

variations of cluster-decorated Penrose tilings. A typical example

is shown in Fig. 4.

The structure of DQC shows atomic layers not only perpendi-

cular to the tenfold axis but also inclined to it. Consequently,

those slightly puckered layers connect the periodic and the

quasiperiodic directions. This coupling constrains the evolution

of quasiperiodic order during QC growth (see Steurer47 and

references therein). It favors strict quasiperiodic order because

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 9: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

6726 Chem. Soc. Rev., 2012, 41, 6719–6729 This journal is c The Royal Society of Chemistry 2012

any deviation from it would hinder the growth of the inclined

atomic layers. Consequently, 2D growth models or 3D models

assuming stackings of largely independent 2D quasiperiodic layers

are not realistic. They would lead to random-tiling-like structures,

indeed. It has been demonstrated by matching-rule-based Monte

Carlo (MC) simulations that a 2D rhomb Penrose tiling, existing

as a ground state at T = 0, transforms more and more into a

random tiling with increasing temperature.48 The phason

fluctuations appear to be unbounded, showing a logarithmic

divergence on increasing system size. This means that at T> 0

in thermodynamic equilibrium randomized Penrose tiling

possesses only quasi-lro. Its Fourier spectrum would not

contain a pure point contribution (Bragg reflections) anymore.

While macroscopic single crystals of Al-based DQC easily

form, this is different for Zn–Mg-based ones. In the system

Zn–Mg–Dy, for instance, Frank–Kasper-type DQC and IQC

exist at slightly differing compositions. While macroscopic

single crystals of the IQC can be quite easily obtained, in the

case of the DQC it took 10 months annealing time atE650 K,

indicating a rather sluggish, diffusion based solid state reaction.

3.2 Dodecagonal quasicrystals (DDQC)

On the atomic scale, there is just one, probably stable,

dodecagonal QC known, dd-Ta1.6Te, with a layer structure

based on a triangle–square tiling.40,41 The layer structure causes

its soft lubricant-like appearance. A detailed structure model

of the fundamental building cluster (Ø E 20.5 A) could be

only derived from its approximant oP628-Ta97Te60.14 The

Ta-substructure is amazingly similar to that of the spherical

assemblies of dendritic LQC (compare Fig. 2 of Conrad and

Harbrecht14 with Fig. 3 of Zeng et al.15). The corrugated

lamellae, each consisting of three Ta-layers sandwiched between

two hcp Te-layers, contain two length-scales: that of strong Ta–Ta

and Ta–Te interactions (dTa–Ta = 2.95 A, dTa–Te = 3.05 A)

and that of weak Te–Te interactions (dTe–Te = 3.34 A). The

Ta-substructure can be seen as tcp structure. The Te-atoms

correspond to the periodic average structure (PAS) of the square–

triangle tiling of the Tac-atoms at the centers of the hexagonal

Ta-antiprisms (see Fig. 3 of Krumeich et al.41). The structure

of the DDQC has been also described as modulated hexagonal

structure with its origin in a charge-density wave (CDW).49

Due to the quite anisotropic layer structure of dd-Ta1.6Te,

this quasiperiodic structure is much closer to a 2D quasi-

periodic system than those of DQC. Consequently, one could

expect a lower degree of quasiperiodic lro. Indeed, this DDQC

is of rather poor quality as indicated in Fig. 1 of Krumeich

et al.41 It shows periodic nanodomains, mainly composed of

square tiles, beside some random-tiling-like ones. Overall, the

structure resembles that of entropy-stabilized random tilings.50

In thermal annealing experiments for several weeks, the

DDQC appeared stable for T = 1070 K and transformed into

approximant (?) structures for T > 1800 K.40 However, there

is not enough experimental information so far to decide on the

thermodynamic stability of dd-Ta1.6Te.

3.3 Icosahedral quasicrystals (IQC)

IQC represent the largest fraction of stable QC. There are

binary and ternary IQC known so far (Fig. 1). Contrary to

2D quasiperiodic systems, phason fluctuations in 3D quasi-

periodic tilings remain bounded with increasing temperature.

This has been demonstrated by MC simulations on a 3D

Ammann tiling.51 Consequently, the thermodynamic equili-

brium randomized 3D Ammann tiling is on average still

quasiperiodic resulting in a Fourier spectrum containing both

Bragg peaks and continuous diffuse scattering.

In an X-ray diffraction study on icosahedral Al–Cu–Fe with

very large dynamic intensity range, it was estimated that

the amplitude of phason fluctuations reaches E7% of the

diameter of the largest atomic surface (perp-space component

of the 6D hyperatoms) and E19% of the smallest one. This

means that more than 20% of all atoms are involved in

phasonic (including chemical) disorder.52 This indicates a

disordered quasiperiodic structure rather than a random-

tiling-like one, where this fraction would be much higher.

4 Mesoscopic (soft) quasicrystals

In recent years an increasing number of soft QC has been

identified with self-assembled structures on the mesoscale.

Examples are structures of supramolecular dendrimer

LQC,15 ABC star terpolymers16 and colloidal micelles in

solution19 (for reviews see the publications of Dotera,20 Ungar

et al.21 and Dotera53). Furthermore, mesoscopic quasicrystals

were also obtained by self-assembly of surfactant-coated

metallic nanoparticles.17 Contrary to hard QC, some soft

QC can self-assemble from just one type of particles.

The dominating factors determining whether a structure

gets periodic or quasiperiodic are the chemical composition

and cluster shape in the case of intermetallic QC. In the case of

colloidal QC, hard-core–soft-shell potentials seem to be deci-

sive while in the case of polymeric QC the shape (terpolymer)

and stoichiometry plays a role again. In all cases, double-well

potentials favoring two length scales (see Fig. 8(e)) appear to

be responsible for the formation of the structural subunits. In

the case of colloidal QC the temperature determines whether a

periodic or quasiperiodic structure forms, in the other cases it

is mainly the stoichometry.

Without an external field, all QC-forming mesoscopic

phases assemble themselves in structures based on either

periodic Archimedean or quasiperiodic dodecagonal tilings.

In both cases, the prototiles are triangles and squares with

frequencies according to the kind of tiling and depending on

composition and/or temperature.

In contrast to hard QC, where this symmetry has never been

observed, micrometre-sized polystyrene particles, dispersed in

water, self-assemble with 7-fold quasiperiodic symmetry if

they are exposed to a weak potential created by interference

of seven Laser beams.18 The authors relate the probability of

the formation of quasiperiodic structures with n-fold symmetry

to the density of high-symmetry stars (local n-fold symmetry),

which decreses with n. Holographic trapping of silica micro-

spheres in aqueous solution was used to assemble QC with 5-,

7- and 8-fold symmetry.54 Another study suggests that the

probability of the existence of a QC with n-fold symmetry

decreases with increasing dimensionality dAS of the atomic

surfaces (AS) of the embedded quasiperiodic structure. For

n=5, 8, 10 and 12 the dimensionality dAS = 2, for icosahedral

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 10: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

This journal is c The Royal Society of Chemistry 2012 Chem. Soc. Rev., 2012, 41, 6719–6729 6727

QC dAS = 3, for n= 7 and 9 dAS = 4, for n= 15 dAS = 6, for

n = 11 and 13 dAS = 8 and 10, respectively. Indeed, there are

also colloidal structures known that order with 9-fold symmetry

in an external field.19

4.1 Three-component systems – ABC star terpolymers

These three-arm star-shaped copolymers consist of three

different immiscible polymers A, B and C of similar molecular

weight linked at a junction point. In the liquid state, phase-

separated columnar nanodomains are formed with the junction

points lining up along their axes, where the three interfaces

between A/B, B/C and C/A meet. The cross sections of the

solidified terpolymers show polygonal structures related to

Archimedean tilings16 (Fig. 8). Depending on the frequencies

of triangles and squares in the skeleton of the tiling, which can

be varied by changing the arm lengths (volume fractions) of

A, B or C, either periodic Archimedean tilings or on average

quasiperiodic, dodecagonal random tilings result. The three

different nanodomains A, B, and C correspond to the regular

polygons in the Archimedean tilings shown in Fig. 8(b–d).

The driving force for the formation of these structures is seen

in the minimization of the interfacial area. Phason fluctuations,

which would lead to a particular reshuffling of the unit tiles,

are kinetically hindered by the prismatic morphology of the

nanodomains.

MC simulations based on the diagonal-bond method revealed

a phase sequence 4.82 - 32.4.3.4 - DDQC - 4.6.12 with

increasing C content.53 Thereby the skeleton connecting the C

nano domains changes from a square tiling (4-fold symmetry)

to a triangle tiling (6-fold symmetry) via square–triangle tilings.

The 32.4.3.4 Archimedean tiling has a ratio r = Nt/Ns = 2 of

the numbers of triangles and squares compared to the ratio for

the dodecagonal tiling r ¼ 4=ffiffiffi3p¼ 2:309. Consequently, the

crucial factor for the formation of the quasiperiodic structure is

the right value for the ratio r. A prerequisite is further the

constraint that only triangular and square arrangements of the

nano domains are possible due to the usage of star terpolymers

with specific interaction potentials, which, however, do not

essentially change by changing the volume fraction of C.

In terms of the phenomenological Landau theory it was

found that the third-order terms in the free energy favor these

kinds of tiling structures.55 This can also be modeled by the

Lifshitz–Petrich equation, which requires two characteristic

length-scales and effective three-body interactions.56,57

4.2 Two-component systems – surfactant-coated metallic

nanoparticles

A large variety of different structures can form by self-assembly

of surfactant-coated nanoparticles depending on their size

ratios and concentration. The surfactant molecules, which

usually have long hydrocarbon chains, prevent aggregations

of the nanoparticles by a soft repulsive potential counter-

balancing van der Waals interactions. In a similar way as that

of soft colloidal quasicrystals (see Section 4.3), they form tcp

structures, periodic as well as quasiperiodic ones. Examples are

mixtures of surfactant-coated nanoparticles of Fe2O3 (Ø 134 A)

and Au (Ø 50 A). As a function of the ratio of these two kinds

of nanoparticles all kinds of structures self-assemble, which

can be described as decorated triangle–square tilings when

connecting the centers of the Fe2O3 nanoparticles:17 from the

triangle tiling 36, via the Archimedean tiling 32.4.3.4 (related to

the Frank–Kasper s phase, see Fig. 9), and the dodecagonal

tiling to the square tiling, 44. Each triangle is centered by one

Au nanoparticle and each square by a cluster of six. For a

ratio of the number of triangles to the number of squares of

N3=N4 ¼ 4=ffiffiffi3p� 2:31, what also means that the area covered

by triangles and squares is equal, the random dodecagonal

tiling has its ideal composition.58

Fig. 8 Self-assembly of ABC-star terpolymers. (a) The polymer

chains, linked together at a single junction point, form a 63 tiling for

equal amounts of the monomers A, B and C. In the case of different

monomer ratios, structures form based on the triangle–square tilings

(b) 4.6.12, (c) 4.82 and (d) 32.4.3.4. A vertex configuration nm is defined

by the kind of polygons arranged along a circuit around a vertex. The

unit cells are outlined. (e) Dodecagonal tiling formed by a hypothetical

ABC-star terpolymer with concentrations of C > A > B. The black

arrows mark the two length scales a and affiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2þ

ffiffiffi3pp¼ 1:93a.

Fig. 9 (a) Structure of the Frank–Kasper phase s-Al40Ta60 projected

along [001] (Al—gray, Ta—black). The centers of the hexagonal

antiprismatic columns form the Archimedean square–triangle tiling

32.4.3.4 (outlined red). (b) Icosahedral AET (blue, and marked also

blue in (a)) appear in the s-phase along the tile edges, where the

hexagonal antiprismatic columns meet. In the case of an idealized

colloidal DDQC, the atoms are replaced by spherical micelles and the

triangular and square unit tiles are arranged like in Fig. 8(e).

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 11: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

6728 Chem. Soc. Rev., 2012, 41, 6719–6729 This journal is c The Royal Society of Chemistry 2012

Since similar structures are also found in other nanoparticle

systems, the authors assume the DDQC to be, for this

composition, the ‘‘most stable, highest entropy quasicrystal

phase of very weakly attractive spherical particles’’.17

MD simulations of systems consisting of micelles with

mobile surface features and a significant fraction of dimers,

with aspect ratio 1.5 for disfavoring close-packed structures,

lead to DDQC as well, even when using simple, purely repulsive

potentials.59 While double-well potentials such as the Dzugutov

potential60 can lead to energy-stabilized DDQC, this technique

yields entropy-stabilized DDQC.

4.3 Single-component systems – colloidal quasicrystals and

hard polyhedra

It has been shown in several experimental and theoretical

studies that colloids in solution can self-assemble into rather

complex structures, from Frank–Kasper phases to dodecagonal

QC. Examples are supramolecular dendrimer LQC15,21 and

colloidal micelles in solution.19 The decisive factors for the

formation of the different structures, periodic or quasiperiodic,

are specific interaction potentials as well as the temperature.

The pair potential of a micelle is characterized by a strongly

repulsive (hard) core, a van-der-Waals attraction at the core

boundary and an exponentially decaying (soft) repulsive inter-

action in the outer part (shell) (see, e.g., Barkan et al.57).

An important feature for the suppression of close packings

is the formation of icosahedra centered at sites where the

hexagonal antiprisms share vertices (Fig. 9).

It has been shown by MD simulations that rather simple

double-well Lennard-Jones–Gauss interaction potentials can

lead to quasiperiodic structures.38 Depending on the locations

of the first and the second minimum, which define two

competing nearest-neighbor distances, pentagonal, decagonal

or dodecagonal structures evolve. The dodecagonal structure

has the highest density of all these quasiperiodic phases. An

alternative mechanism is solely based on particle functionali-

zation and shape in order to disfavor all kinds of close

packings, encouraging the formation of structures with low

surface contact area.59

The formation of periodic or quasiperiodic tcp structures

that are known from transition metal compounds is explained

by the hard-core–soft-shell potential of the colloidal particle.

In a similar way, transition metals with their slowly decaying d

and f orbitals (less-localized electrons than in s or p orbitals)

can be seen as spheres with a soft shell. Such spheres energe-

tically prefer tcp structures over other close-packed structures,

which have octahedral voids in addition to the much smaller

tetrahedral ones.21

It has been shown in several model simulations that entropy

driven, on-average dodecahedral quasiperiodic ordering can

result under certain conditions for hard complex objects such

as tetrahedra or dimers of tetrahedra.29,30 In these cases, the

two length scales are provided by the shape of the particles,

which also disfavors cubic or hexagonal close packings.

5 Summary

What are the ingredients needed for the formation of quasi-

periodic structures in both atomic and mesoscopic systems?

First of all, the fundamental building units, atoms, micelles,

surfactant-coated nanoparticles or columnar nanodomains

must have interaction potentials forcing the formation of

particular structural subunits such as clusters or patches of

unit tiles. Model calculations have shown that particularly

suitable can be double-well potentials defining two different

length-scales characteristic for particular quasiperiodic structures.

In the case of intermetallic QC, Friedel-oscillations in the pair

potentials fulfill this criterion, in the case of mesoscopic QC, they

result from the hard-core–soft-shell structure of the building units.

Whether these energetically favorable structural units are

packed periodically or quasiperiodically depends on the following

factors: their shape and the relationship between their chemical

composition and that of the whole phase as well as the electronic

band structure in the case of hard QC, while it is mainly

composition and temperature in the case of soft QC.

In both cases, entropic contributions from phason fluctuations

and/or other kinds of disorder can be crucial for the stability of

the quasiperiodic phase. The much weaker interactions on the

mesoscopic scale increase the weight of entropic contributions

leading to random(ized) tiling based structures. The prevalence

of 12-fold symmetry in soft QC results from the preference

of their soft-shell particles for tetrahedrally-close-packings.

A somehow similar behavior shows the hard QC dd-Ta1.6Te,

with the substructure of ‘‘soft’’ Ta atoms close to a tcp

packing. For all other hard QC, the 5-fold symmetry of their

fundamental building clusters forces the formation of QC with

5-fold symmetry axes (DQC, IQC).

Finally, I want to emphasize that the origin of quasi-

periodicity is already much better understood for mesoscopic

QC than for intermetallic QC. What is still needed to under-

stand QC formation and growth on the atomic scale is a growth

simulation for intermetallic QC under realistic conditions, i.e.,

with large model system sizes (up to millions of atoms) and the

full atom dynamics at the high temperature of formation.

References

1 D. Levine and P. J. Steinhardt, Phys. Rev. Lett., 1984, 53, 2477–2480.2 D. Shechtman, I. Blech, D. Gratias and J. W. Cahn, Phys. Rev.Lett., 1984, 53, 1951–1953.

3 R. J. Schaefer, L. A. Bendersky, D. Shechtman, W. J. Boettingerand F. S. Biancaniello, Metall. Trans. A, 1986, 17, 2117–2125.

4 K. F. Kelton, J. Non-Cryst. Solids, 2004, 334, 253–258.5 D. Holland-Moritz, T. Schenk, V. Simonet and R. Bellissent,Philos. Mag., 2006, 86, 255–262.

6 O. S. Roik, S. M. Galushko, O. V. Samsonnikov, V. P. Kazimirovand V. E. Sokolskii, J. Non-Cryst. Solids, 2011, 357, 1147–1152.

7 W. Steurer, Philos. Mag., 2006, 86, 1105–1113.8 C. L. Henley, M. de Boissieu and W. Steurer, Philos. Mag., 2006,86, 1131–1151.

9 P. J. Steinhardt, H. C. Jeong, K. Saitoh, M. Tanaka, E. Abe andA. P. Tsai, Nature, 1998, 396, 55–57.

10 P. Saintfort, B. Dubost and A. Dubus, C. R. Seances Acad. Sci.,Ser. 2, 1985, 301, 689–692.

11 A. P. Tsai, Acc. Chem. Res., 2003, 36, 31–38.12 W. Steurer and S. Deloudi, Crystallography of Quasicrystals –

Concepts, Methods and Structures, Springer, Berlin, Heidelberg,2009, vol. 126.

13 Y. Miyazaki, J. Okada, E. Abe, Y. Yokoyama and K. Kimura,J. Phys. Soc. Jpn., 2010, 79, 073601.

14 M. Conrad and B. Harbrecht, Chem.–Eur. J., 2002, 8, 3094–3102.15 X. B. Zeng, G. Ungar, Y. S. Liu, V. Percec, S. E. Dulcey and

J. K. Hobbs, Nature, 2004, 428, 157–160.

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online

Page 12: Chemical Society Reviews - stuba.skszolcsanyi/education/files/The Nobel...the determination of their structures and the understanding of their physical properties. Now, quasiperiodic

This journal is c The Royal Society of Chemistry 2012 Chem. Soc. Rev., 2012, 41, 6719–6729 6729

16 K. Hayashida, T. Dotera, A. Takano and Y. Matsushita, Phys.Rev. Lett., 2007, 98, 195502.

17 D. V. Talapin, E. V. Shevchenko, M. I. Bodnarchuk, X. C. Ye,J. Chen and C. B. Murray, Nature, 2009, 461, 964–967.

18 J. Mikhael, M. Schmiedeberg, S. Rausch, J. Roth, H. Stark andC. Bechinger, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 7214–7218.

19 S. Fischer, A. Exner, K. Zielske, J. Perlich, S. Deloudi, W. Steurer,P. Lindner and S. Forster, Proc. Natl. Acad. Sci. U. S. A., 2011,108, 1810–1814.

20 T. Dotera, Isr. J. Chem., 2011, 51, 1197–1205.21 G. Ungar, V. Perec, X. Zeng and P. Leowanawat, Isr. J. Chem.,

2011, 51, 1206–1215.22 R. Penrose, in Tilings and Quasicrystals: a non-local growth problem?

ed. M. V. Jaric, Academic Press Inc (London) Ltd, 1989, vol. 3, ch. 2,pp. 53–79.

23 T. Janssen, G. Chapuis and M. de Boissieu, Aperiodic Crystals.From Modulated Phases to Quasicrystals, Oxford University Press,2007, vol. 20.

24 U. Mizutani, Hume-Rothery Rules for Structurally Complex AlloyPhases, CRC Press, 2011.

25 J. Dshemuchadse, D. Jung and W. Steurer, Acta Crystallogr., SectB: Struct. Sci., 2011, 67, 269–292.

26 C. Gomez and S. Lidin,Angew. Chem., Int. Ed., 2001, 40, 4037–4039.27 G. Coddens and W. Steurer, Phys. Rev. B: Condens. Matter Mater.

Phys., 1999, 60, 270–276.28 W. Steurer and S. Deloudi, Struct. Chem., 2012, DOI: 10.1007/

s11224-011-9864-2.29 A. Haji-Akbari, M. Engel, A. S. Keys, X. Y. Zheng,

R. G. Petschek, P. Palffy-Muhoray and S. C. Glotzer, Nature,2009, 462, 773.

30 A. Haji-Akbari, M. Engel and S. C. Glotzer, Phys. Rev. Lett.,2011, 107, 5702.

31 W. Steurer, Z. Kristallogr., 2004, 219, 391–446.32 W. Steurer and S. Deloudi, Acta Crystallogr., Sect. A: Found.

Crystallogr., 2008, 64, 1–11.33 P. Gummelt, Geom. Dedic., 1996, 62, 1–17.34 P. Villars andK. Cenzual,Pearson’s Crystal Data, ASM International,

Ohio, USA, 2011/2012.

35 T. Weber, J. Dshemuchadse, M. Kobas, M. Conrad, B. Harbrechtand W. Steurer, Acta Crystallogr., Sect. B: Struct. Sci., 2009, 65,308–317.

36 W. Steurer, Z. Kristallogr., 2000, 215, 323–334.37 M. Widom, I. Al-Lehyani and M. Mihalkovic, J. Non-Cryst.

Solids, 2004, 334, 86–90.38 M. Engel and H. R. Trebin, Phys. Rev. Lett., 2007, 98, 225505.39 D. Schopf, P. Brommer, B. Frigan and H. R. Trebin, Phys. Rev. B:

Condens. Matter Mater. Phys. (volume, page), 2012.40 M. Conrad, F. Krumeich and B. Harbrecht, Angew. Chem., Int. Ed.,

1998, 37, 1384–1386.41 F. Krumeich, M. Conrad, H. U. Nissen and B. Harbrecht, Philos.

Mag. Lett., 1998, 78, 357–367.42 M. Widom and E. Cockayne, Physica A, 1996, 232, 713–722.43 J. Roth and C. L. Henley, Philos. Mag. A, 1997, 75, 861–887.44 W. Steurer, Philos. Mag., 2007, 87, 2707–2712.45 H. Orsini-Rosenberg and W. Steurer, Philos. Mag., 2011, 91,

2567–2578.46 P. Gummelt, J. Non-Cryst. Solids, 2004, 334, 62–67.47 W. Steurer, Z. Anorg. Allg. Chem., 2011, 637, 1943–1947.48 L. H. Tang and M. V. Jaric, Phys. Rev. B: Condens. Matter Mater.

Phys., 1990, 41, 4524–4546.49 M. Uchida and S. Horiuchi, J. Appl. Crystallogr., 1998, 31, 634–637.50 D. Joseph and V. Elser, Phys. Rev. Lett., 1997, 79, 1066–1069.51 L. H. Tang, Phys. Rev. Lett., 1990, 64, 2390–2393.52 T. Weber, S. Deloudi, M. Kobas, Y. Yokoyama, A. Inoue and

W. Steurer, J. Appl. Crystallogr., 2008, 41, 669–674.53 T. Dotera, J. Polym. Sci. B: Polym. Phys., 2012, 50, 155–167.54 Y. Roichman and D. G. Grier, Opt. Express, 2005, 13, 5434–5439.55 T. Dotera, Philos. Mag., 2007, 87, 3011–3019.56 R. Lifshitz and H. Diamant, Philos. Mag., 2007, 87, 3021–3030.57 K. Barkan, H. Diamant and R. Lifshitz, Phys. Rev. B: Condens.

Matter Mater. Phys., 2011, 83, 172201.58 P. W. Leung, C. L. Henley and G. V. Chester, Phys. Rev. B:

Condens. Matter Mater. Phys., 1989, 39, 446–458.59 C. R. Iacovella, A. S. Keys and S. C. Glotzer, Proc. Natl. Acad.

Sci. U. S. A., 2011, 108, 20935–20940.60 M. Dzugutov, Phys. Rev. A, 1992, 46, R2984–R2987.

Dow

nloa

ded

on 2

7 Se

ptem

ber

2012

Publ

ishe

d on

22

May

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2CS3

5063

G

View Online