Mathematics of Quasi-Crystals
Mathematics of Quasi-Crystals
Michael Baake
Institut für Theoretische Physik, Universität Tübingen,
Auf der Morgenstelle 14, D-72076 Tübingen, Germany
1. Introduction
arXiv:math-ph/9901014v1 20 Jan 1999
The discovery of alloys with long-range orientational order and sharp diffrac-
tion images of non-crystallographic symmetry [65, 35] has initiated an inten-
sive investigation of the possible structures and physical properties of such
systems. Although there were various precursors, both theoretically and ex-
perimentally [72], it was this renewed and amplified interest that established
a new branch of solid state physics, and also of discrete geometry. It is usu-
ally called the theory of quasicrystals, even though it also covers ordered
structures more general than those with pure Bragg diffraction spectrum.
It is now rather common to think of the regime between crystallographic
and amorphous systems as an interesting area with a hierarchy of ordered
states. This was not so some fifteen years ago, and it is the purpose of this
contribution, and of the book as a whole, to introduce some of the ideas
and methods that are needed to handle this new zoo. In particular, I will
summarize some mathematical and conceptual issues connected with it, with
special emphasis on proper equivalence concepts. This is more important
than it might appear at first sight, because non-periodic order shows both
new features and new hazards – and it is worthless to talk about a property of
one specific structure if it is lost for others that are locally indistinguishable.
To develop some of these ideas, one has to start with a valid idealization
of the physical structures one has in mind. Since we are interested in solids
of some relevant size here, it is reasonable to replace their atomic arrange-
ments by suitable infinite point sets. These should be uniformly discrete (i.e.,
there should be a uniform minimal distance between the points) and, usually,
they should be relatively dense (i.e., there is a maximal hole). Sets with this
property are called Delone sets and are widely used for this purpose.
In analogy to ordinary crystallography, many people prefer to think in
terms of cells or tiles [56]. Here, one may start from a (usually finite) number
of proto-tiles that fit together to tile space without gaps or overlaps. If we now
decorate the tiles by finitely many points (giving the atomic positions, say),
we return to a Delone set. Vice versa, given a Delone set Λ, we can perform
the Voronoi construction that attaches to each point x ∈ Λ the region of
all points of ambient space that are closer to x than to any other point of
Λ. This way, we come back to a tiling (whose dual, the so-called Delone
tiling [66] is an even better candidate). Under an additional, but rather mild,
2 Michael Baake
2. Non-crystallographic symmetries
Section) and then why their appearance in the diffraction images of solids is,
at least at first sight, astonishing (to be addressed in the next Section).
3. Diffraction
To simplify things, we will only talk about kinematic diffraction, i.e. diffrac-
tion that can be understood in terms of single scattering in the Fraunhofer
picture. This is quite appropriate for X-ray and neutron diffraction, but not
for electron diffraction where multiple scattering is essential, see [18] for de-
tails. Kinematic diffraction from a structure, in turn, is closely related to the
Fourier transform of the corresponding potential in the sense that the ob-
served intensities of sharp spots (Bragg peaks) are proportional to the abso-
lute squares of the Fourier amplitudes, see [18] for a more detailed discussion
and justification of this point of view.
In our idealized world, with atoms etc. replaced by point scatterers (Dirac
distributions) on a set Λ, we thus consider the so-called Dirac comb of Λ,
X
ω = ωΛ := δx , (3.1)
x∈Λ
where δx is Dirac’s distribution (or measure) at point x, i.e. (δx , ψ) = ψ(x) for
all test functions ψ(x). In a second step, one defines a so-called autocorrelation
or Patterson function [18] for this,
1 X
γω := lim δx−y , (3.2)
r→∞ vol(Br (0))
x,y∈Λ∩Br (0)
A Guide to Mathematical Quasicrystals 5
Fig. 2.2. Diffraction image of the Ammann-Beenker tiling of Figure 2.1. The area
of each disc is proportional to the intensity of the corresponding peak, the cut-off
is at 0.1 % of the central intensity.
where Br (0) is the (solid) ball of radius r around 0, and the limit is assumed
to exist. Furthermore, we will tacitly assume that this limit stays the same if
we replace the ball by any other convex region centred around the origin (with
r the radius of the maximal inscribed ball) – an actually rather non-trivial
feature to establish. The autocorrelation is a distribution of the form
X
γω = ν(z)δz (3.3)
z∈∆
ˇ
With this convention, one has ǧˆ = g and fˆ = f . Also, there is no need to
distinguish between dual and reciprocal lattice (the factor 2π is absorbed into
6 Michael Baake
the argument of the exponential function) and the convolution theorem takes
the nice form f[
∗ g = fˆ · ĝ where
Z
(f ∗ g)(x) = f (x − y)g(y)dy . (3.5)
Rn
FT ↓ ↓ FT (3.9)
|.|2
ω̂ −→ γ̂ω
Whenever this situation applies, in the sense that all quantities exist and the
diagram is commutative, things are rather simple. In particular, given the sit-
uation of a lattice, atomic profiles (extended scatterers) or more complicated
decorations of a fundamental domain (multiple atoms per unit cell) can be
incorporated by means of convolutions with ω and then be processed through
the Wiener diagram. Formally, the same process is then always used (at least
for the Bragg part), but this needs extra justification, and often hard analysis
for a proof, compare [33, 70] and references therein for details.
A Guide to Mathematical Quasicrystals 7
It is now time to explain how such a strange diffraction behaviour can come
along. The foundation of it goes back to the beginning of the century when
Harald Bohr, the younger brother of Niels, developed what is now called the
theory of almost periodic functions. Let us consider the example
2
We only talk about diffraction from “single crystals” here, resp. its analogue for
quasicrystals. A somewhat similar discussion appears for the diffraction from
powders, but should be distinguished clearly, because this case has a rather
different explanation, see chapter 16 of [18].
8 Michael Baake
√
where τ = (1 + 5)/2 is the famous golden ratio. This is an irrational number
(in fact, as follows from its continued fraction expansion, the most irrational
one), whence f (x) is certainly not periodic. Nevertheless, for any given ε > 0,
there are “almost-translations” t such that |f (x)− f (x+ t)| < ε, for all x ∈ R.
Furthermore, such translations are not rare, but lie relatively dense in R, i.e.
there is a maximal distance between any two consecutive ones. The set of
continuous functions with this property is closed under uniform convergence,
and can be uniformly approximated by trigonometric polynomials. This re-
sults in a generalization of Fourier series which is essentially the core of Bohr’s
work [15]. For a more recent introduction, with additional material, see [17].
In these generalized Fourier series, pairwise incommensurate base frequen-
cies occur (such as 1 and τ in the above example). If their number is finite,
the corresponding function is called quasiperiodic. This subclass of functions
has the property that it can be obtained as a section through a periodic
function of more variables, e.g., in our example,
f (x) = sin(x) + sin(y)|y=τ x (4.2)
This is also the essential idea to understand the diffractivity of quasicrystals,
see [44, 23, 40] and various articles in [71].
Let us therefore construct non-periodic point sets by suitable sections
through a crystallographic structure in higher dimension. As a first step,
A Guide to Mathematical Quasicrystals 9
let us take a look at the so-called cut and project method, an example of
which is shown in Figure 4.1. Starting with the square lattice in the plane,
Z2 , a line with irrational slope, called E, is drawn, surrounded by a parallel
strip of finite width. All lattice points inside the strip are then projected to
E. The result is a sequence of points that forms a non-periodic Delone set
(due to the irrationality of the slope – otherwise it would be periodic). If
the slope (as in Figure 4.1) is 1/τ , and if the width of the strip coincides
with the projection of a fundamental square to the internal direction (which
is actually perpendicular here), Eint , we obtain what is called the Fibonacci
chain, the most common and best studied non-periodic 1D point set. Note
that we have not given a formal definition of a quasicrystal3, and we will not
do so because the present use of the word is far from being context-free, and
a really natural approach is not yet in sight. Let us add that some authors
would prefer not to call the Fibonacci chain a real quasicrystal, but rather
a modulated crystal. The reason is topological in nature, compare [38], but
since such aspects are not important in our present context, we will suppress
them.
This projection scheme, which has an obvious generalization in higher
dimensions, does not seem to be an exact analogue of the section idea men-
tioned before, but it is equivalent to it. To see this, take the intersection of
the strip with Eint , which is an interval here. This set, W , is called window
or acceptance domain, and our Fibonacci chain F is then given by
F = {P (x) | x ∈ Z2 and Pint (x) ∈ W } , (4.3)
where P and Pint denote the canonical projections to E and Eint , respectively4 .
The same set is obtained if we, instead of using the strip method, take an
inverted copy of the window, −W , stitch it to each lattice point of Z2 , and
modify the rule in saying that we get a point of F whenever our cut line, E,
crosses a copy of this set. It can be considered as a target or a kind of atomic
hypersurface which is point-like in the direction of the “physical” space, E,
and extended only in “internal” space, Eint . A third method to describe the
same object goes under the name dualization scheme and has the advantage
of directly giving cells rather than point sets, see [57, 46, 66, 9] for details.
In view of the Fourier transform, the version with the atomic hypersur-
faces seems most attractive, because it is closest to the idea of describing
a quasiperiodic arrangement of scatterers as a section through a crystallo-
graphic arrangement in higher dimensions. Consider now a Fibonacci chain,
F , with point scatterers of equal strength on all its points, i.e. consider the
Dirac comb X
ωF = δx . (4.4)
x∈F
3
A reasonable working definition of a quasicrystal would include all discrete
patterns which possess an autocorrelation whose Fourier transform (i.e. the
diffraction) is either purely discrete, or has at least a non-trivial discrete part.
4
For a more general and systematic formulation, we refer to [54, 67, 66].
10 Michael Baake
Eint
E
(0,1)
(1,0)
Fig. 4.1. Projection method for the Fibonacci chain and torus parametrization of
its LI-class.
By a simple (formal) calculation, one finds that the Fourier transform consists
of Dirac peaks on all points P (k) where k is a point of the dual of the
embedding lattice. Its amplitude, a(P (k)), is formally given by
Z
d
a(P (k)) = e−2πi kint xint dxint (4.5)
vol(W ) −W
where xint = Pint (x) etc. and d = dens(F ) denotes the density of F . The
diffraction image is composed of Bragg peaks at the points P (k) of intensity
|a(P (k))|2 . The derivation of this can be found in many articles, but has to be
taken with a grain of salt: it is purely formal, because the resulting expression
is not a locally summable distribution and hence not a valid representation
of a tempered distribution. That this formal way of calculating amplitudes
and intensities is nevertheless correct, was proved much later by Hof, and the
interested reader is referred to [33] and references therein.
In what sense does all this resolve the puzzle we started from? If we take
a closer look at Eq. (4.5), we realize that, if W is an interval, the abso-
lute squares of the amplitudes are of the form sin(2πkint )2 /(2πkint )2 , hence
bounded by c/|kint |2 with some constant c – and this means that only finitely
A Guide to Mathematical Quasicrystals 11
many peaks per unit volume have an intensity beyond a given threshold be-
cause |kint | is the distance of the (dual) lattice point k from the cut space. A
cut-off for the intensities thus has an effect similar to the projection method
itself! This is perhaps one of the most important observations in this context:
a point set can be diffractive, and show a clear signature of this, without be-
ing crystallographic. If this is the case, there is then no longer any reason why
non-crystallographic symmetries should not show up. If they do, however, we
know immediately that the system cannot be crystallographic, and we would
try to use the idea of a section through a lattice in higher dimensions to
describe the structure.
Let me close this Section with another warning. The success of the projec-
tion method does not indicate that there is any need for higher-dimensional
physics. It is only a convenient description of a certain class of ordered struc-
tures. Clearly, it is tempting to derive all sorts of generalizations of common
properties and theorems (e.g. Bloch’s theorem) by employing the embedding
scheme and a chain of formal calculations. Quite frequently, this leads to
wrong conclusions, and extreme care is required. For example, there is no easy
analogue of Bloch’s theorem. In fact, its naive generalization fails as badly
as possible: the standard tight-binding model on the Fibonacci chain, in the
infinite size limit, has no bands at all, and the spectrum is neither absolutely
continuous nor pure point, but purely singular continuous! In other words,
it is precisely of the form that was argued impossible for physical structures
not too long ago. For more on this, and on the existence of a Cantor-type
gap structure with topological quantum numbers, see [73, 13, 2].
Let us note that the minimal dimension is usually sufficient (unless one
wants to describe “modulated” quasicrystals, where it doubles), and using
more than the minimal number only results in ambiguities of the indexing
scheme – an altogether undesired feature.
Having settled the question for the correct dimension, we need to know
what the “right” lattice is. It turns out that the higher-dimensional analogue
of the square and cubic lattices, the hypercubic lattices, are not sufficient.
The most common example where this becomes apparent is the Penrose tiling
of Figure 3.1. It has fivefold (actually tenfold) symmetry6 , and the above
Theorem then tells us that a 4D lattice is the right choice, because φ(5) =
φ(10) = 4. Very often, one finds a description of the Penrose tiling based upon
Z5 where one extra dimension has been introduced. This has the disadvantage
mentioned. A simpler choice is the so-called root lattice A4 which can be seen
as the 4D lattice that is obtained by intersecting Z5 with the 4D hyperplane
through the origin, and orthogonal to the space diagonal (1, 1, 1, 1, 1). In
general, root lattices provide a very nice class of simple lattices that is general
enough to cover the observed cases [7] in a maximally symmetric way. For
background material on root lattices, and all sorts of interesting connections
to other branches of mathematics, we refer to the bible, [16].
Let us briefly mention some other planar examples. The Ammann-Beenker
tiling of Figure 2.1 shows eightfold symmetry, and requires a 4D lattice
(φ(8) = 4). The standard choice [4] is Z4 , but also the face-centred lattice
in 4D is possible, i.e. the root lattice D4 . The latter has the advantage that,
with a different choice of the cut space, also patterns with 12-fold symmetry
(φ(12) = 4, once more) can be obtained, see [5] for details. Sometimes, 12-fold
symmetry is easier to describe with another root lattice, namely A2 × A2 .
Most prominent, in this context, are tilings made from squares and equi-
lateral triangles, such as that shown in Figure 5.1. It is compatible with 12-
fold symmetry (the triangles cover half the area), and was obtained by the
projection method. Its window, however, shows a more complicated struc-
ture: it is a 12-fold symmetric region, compact, the closure of its interior,
but has a fractal boundary, see Figure 5.2. It is a well-accepted conjecture
that all square-triangle tilings with 12-fold symmetry, obtained by projection,
require a fractally shaped window, and, in a certain sense, the one of Fig-
ure 5.1 is an example with “maximally smooth” window boundary [6]: almost
everywhere, the boundary is locally smooth (a line segment, in fact), but at
an uncountable set of boundary points (of vanishing Lebesgue measure) the
fractal dimension is non-integral, and rather close to 2.
After these planar examples, let us briefly sketch the situation in 3D.
Clearly, there are the so-called T-phases (with T for “thumbtack”, to mimic
their geometric structure) which are quasiperiodic in a plane and periodic
along the perpendicular line. They can be modelled by 3D tilings that are
stacked layers, each single layer being made from prisms (as tiles) whose base
6
The proper use of the term symmetry will be explained in the next Section.
A Guide to Mathematical Quasicrystals 13
pattern forms one of the classic planar tilings. Clearly, their Fourier image
needs one extra Miller index, i.e. the Bragg peaks of standard decagonal T-
phases are indexed by 5 integers, 4 being needed for the non-periodic planar
degrees of freedom and one extra index for the periodic direction.
Of greatest importance probably are the tilings with icosahedral symme-
try. Here, one has to distinguish three different types. All can be obtained by
the projection method from hypercubic lattices in 6-space. There are three
different Bravais types of them, the primitive (Z6 ), the face-centred (D6 ),
and the body-centred (D6∗ ) one [68]. The three different icosahedral classes,
see [22] for a detailed description, are then also called primitive (or P -type),
face-centred (F -type) and body-centred (B-type), respectively. Since no ap-
plication of the B-type models are presently known, I’ll skip details of them.
The standard P -type tiling is made from two rhombohedra, an acute
and an obtuse one. It has, in various degrees of completeness, a long history
[72], and was first described by means of the projection technique in [44].
It will be denoted by KN. The diffraction shows icosahedral symmetry and
a clear scaling with inflation multiplier τ 3 . This is characteristic of P -type
14 Michael Baake
structures, and makes the distinction from F -type rather simple as the latter
displays scaling with an inflation multiplier7 τ (the same would be true of
the B-type).
The more important class (in terms of applications) is that of F -type
tilings. One of the earliest examples is the zonohedral tiling by Socolar and
Steinhardt [69], abbreviated as SS. It is built from four proto-tiles, namely
the acute rhombohedron met above, the rhombic dodecahedron, the rhombic
icosahedron and the famous triacontahedron, also known as Kepler’s body
[29]. Another example was found by Danzer [19], which is very closely related
(i.e. locally equivalent) as we shall see later. Danzer’s tiling (called DT) is
built from 4 tetrahedra. Finally, based on the projection technique, several
other F -type tilings have been investigated, see [8, 45] and references therein
for details. The most important of those (called T ∗(2F ) ) is built from six
tetrahedra and is again very closely related to the two tilings mentioned
before (SS and DT), although it contains more local information – a concept
to be made more precise in a shortwhile.
Up to this point, no further mathematical details or concepts were needed
to get a first impression (see [59] for a general construction scheme). But for
a better understanding of the structures, their symmetries and some of the
7
The meaning of this will become clear in the Section on inflation symmetry.
A Guide to Mathematical Quasicrystals 15
new features, we now have to dive a little deeper into the world of discrete
geometry. In particular, we definitely need some good tools to handle the zoo
of possibilities. Later, we shall see that the number of “known” examples, al
least those with “nice” properties, is actually rather small, and can be handled
with little more difficulty than needed for the crystallographic patterns.
One basic concept for the general analysis of global order properties of dis-
crete structures is the equivalence concept of local indistinguishability, also
known as local isomorphism8 [49]. Since the infinite (mathematical) struc-
ture is considered as an approximation to large but finite physical objects, it
is natural to identify those structures which are locally indistinguishable on
arbitrarily large but finite scales. Such structures are called locally indistin-
guishable, or locally isomorphic.
We will use this term frequently in the sequel, so for a precise definition
we introduce some notation. The mathematical objects we deal with are,
most generally, discrete structures in Euclidean space, i.e., sets of (possibly
decorated) bounded subsets of the space which are locally finite in the sense
that each ball of finite radius meets only finitely many structure elements. If
A is such a discrete structure, then we call an r-patch of A each subset of A
which is completely contained in a ball of radius r. Now, two structures, A
and B, are locally indistinguishable (or locally isomorphic) if each r-patch of
A is, up to a translation, also an r-patch of B and vice versa (a moment’s
reflection reveals that, in the general case, this ‘vice versa’ is necessary to
get a proper equivalence relation). The corresponding equivalence class of a
structure is called its local indistinguishability class, or LI-class, for short.
It should be emphasized that this formal definition does not quite reflect
the intuitive description of the first paragraph, because we have insisted on
identity of r-patches up to translations only, rather than up to more general
Euclidean motions. This more restrictive relation will prove useful for other
concepts to be introduced in the next Section, and especially for the definition
of generalized point symmetries.
One of the most outstanding properties of the experimentally observed
aperiodically ordered structures like quasicrystals is the occurence of crystal-
lographically forbidden symmetries in their diffraction spectra, e.g., fivefold
axes. On the other hand, it is clear that, e.g., a 3D discrete structure can
possess at most one axis of exact fivefold point symmetry in a given direc-
tion, because otherwise there would be a dense set of such axes, which is
impossible for a locally finite structure. Therefore, to take into consideration
8
Since this term is occupied with a different meaning in other (connected) areas,
I suggest to replace the word isomorphism by indistinguishability here, which
also avoids the introduction of a new abbreviation.
16 Michael Baake
7. Parametrization of LI-classes
Having given the definition of an LI-class is not quite the same as under-
standing its structure. The latter is, in fact, more complicated than one might
expect. To see this, let us first consider the case of a crystallographic pattern
P: its LI-class consists of all its translates, and can thus be parametrized
by the points of a fundamental domain of the corresponding lattice, Γ , of
translations (e.g., its Voronoi or Wigner-Seitz cell, to be specific) because
P = P + t for all t ∈ Γ . In particular, the LI-class LI(P) simply consists of
one translation class of patterns.
The correspondence between patterns in LI(P) and points of a funda-
mental domain is called the torus parametrization of LI(P) because such a
domain, upon identifying Γ -equivalent boundary points, becomes a torus of
the dimension of the lattice. Clearly, the answer cannot be this simple for non-
crystallographic patterns. Here, LI(P) does not only contain all translates of
P, but also all other patterns that can be obtained as limits of these (w.r.t.
the obvious topology of patch-wise comparison) – and “most” members are
of the latter type. In fact, for repetitive10 aperiodic patterns, the LI-class
9
The symbol Dn appears in two different meanings, once for the corresponding
root lattice and once for the dihedral group of order 2n. Since both are standard
in the literature, and misunderstandings unlikely, we stick to this convention.
10
The term repetitive means the following: for each radius r, there is another
radius, R = R(r), such that each P-patch of radius r can be found in every
P-patch of radius R.
A Guide to Mathematical Quasicrystals 17
11
For certain applications, it is advantageous to distinguish regular and singular
patterns and to adopt a topological point of view, e.g. for questions such as the
spectra of Schrödinger operators, see [13] for details.
18 Michael Baake
the singular members behave, but they are usually less important or even
irrelevant for considerations such as symmetry, inflation invariance etc. In
particular, the torus parametrization allows us to find all members of an LI-
class showing exact invariance under a given symmetry operation, including
new types of symmetry such as inflation/deflation symmetry to be discussed
later. This is based on lifting the symmetry operation under consideration to a
mapping on the torus. Then, the number of fixed points can be determined by
calculating certain determinants. For details, together with explicit examples
and a full treatment of the physically relevant symmetries, see [3, 31].
An extension of this analysis to groups of transformations (rather than
single operations) is possible, and it is instructive to look at subgroups of
the icosahedral group and their action on the three possible types of LI-
classes, see Table 7.1. In each case, there are precisely 64 inversion symmetric
members of the LI-class, and they distribute in a very peculiar way on the
subgroups of Yh , the full icosahedral group. To be more specific, there are two
rhombohedral tilings in LI(KN) with full Yh symmetry, one being regular and
one singular, while there are 4 such members in LI(DT), say, three regular
and one singular. This shows at least one reason why F -type structures are
more frequent than the other possibilities: as a consequence of this analysis,
and using the implications of local indistinguishability, it must be concluded
that F -type tilings or Delone sets have a denser distribution of clusters with
exact (or almost exact) icosahedral symmetry – an idea pretty close to the
concept of a Frank-Kasper phase. This would suggest that F -type icosahedral
quasicrystals should be more frequent than P -type ones, as is indeed the case.
Let us put this type of equivalence in more formal terms. It is clear that the
details of the transformation process described above are not important, e.g.,
20 Michael Baake
we will certainly not care about tiles being properly dissected or composed,
or in fact about tiles at all: quite frequently a representative discrete point
set [42] is what one really needs or wants! The essential feature which allows
the abstraction from local details while keeping track of the global order is
the uniform locality of the transformation rule. It is easy to see that such
a uniformly local rule for the transformation of some structure A into a
structure B exists precisely under the following condition: There is a fixed
finite radius r such that if the r-patches of A around two points, p, q, are
equal up to the translation t = p − q, then the structure B at the points p
and q is the same, again up to the translation t.
If this condition is fulfilled, then we call B locally derivable [11] from A.
If it is also fulfilled with the roles of A and B interchanged (with a possibly
different radius r′ ) then we call A and B mutually locally derivable from one
another, or locally equivalent.
It is clear that this equivalence relation can be extended to entire LI-
classes. That is, if A and B are locally equivalent, then, for any A′ in the
LI-class of A, some B ′ can be found in a canonical fashion (just using “the
same rule”) such that A′ and B ′ are locally equivalent, thereby defining
a one-to-one correspondence between the two LI-classes. Therefore, we can
combine these equivalences defining the MLD-class of a structure A to be the
set of all structures which are locally indistinguishable from some structure
locally equivalent to A. Needless to say that PT and RD in the example above
belong to the same MLD-class in this sense. It is sometimes more useful to
view MLD-classes, which are defined as unions of LI-classes, directly as sets
of LI-classes, but we will identify these two points of view for simplicity.
Let A and B be locally equivalent structures. Obviously, if A is invariant
under a certain translation t, then B must be invariant under t as well, by
the very definition of local equivalence. A little further reflection shows that,
on the other hand, if A and B are crystallographic with the same translation
lattice, Γ , then they are locally equivalent, by “transformation rules” involv-
ing only a couple of fundamental domains. Therefore, local equivalence is a
generalization of “having the same translation lattice” in the periodic case:
Theorem 8.1. Two crystallographic patterns, A and B, are locally equiva-
lent if and only if they share the same translation lattice Γ .
This result also explains why we restricted our definition of local indistin-
guishability (or patch-equivalence) to translations only, rather than using a
version involving congruence. Below, we will refine the MLD concept in order
to achieve a generalization of the space group classification.
formalism and local derivability, see [11]). This way, one can actually prove
that PT and TTT (the so-called Tübingen triangle tiling, another decagonal
tiling built from the golden triangles, see Figure 9.1 and Ref. [9]) are not
in the same MLD-class (although it is possible to rescale PT such that it
becomes locally derivable from TTT). These two LI-classes actually differ in
the distribution of singular tilings, and the transformation rule from TTT to
PT maps certain sets of singular tilings of LI(TTT) onto single, but regular
members of LI(PT) – something that clearly cannot be inverted. This is
remarkable as TTT and PT certainly have the same space group according to
[62], i.e. they cannot be distinguished on the basis of the symmetry properties
of their Fourier transforms.
Fig. 9.1. Decagonal patch of the Tübingen triangle tiling TTT (left) versus Robin-
son’s decomposition (RD) of the rhombic Penrose tiling (right).
Let us now investigate how the MLD-concept works in the context of two
outstanding properties which important aperiodic structures exhibit: infla-
tion/deflation symmetry and perfect matching rules (compare [39, 48, 50, 52,
41] for commonly used definitions).
As there are various concepts of inflation/deflation in the literature, we
have to make precise what we mean by it, thereby taking the opportunity
to put the MLD concept into operation. Usually, an inflation of a structure
consists in a certain rule for a local transformation of structure elements into
patches of a new structure which turns out to be of the same type as the
original one, but on a smaller scale. For example, the dissection of the golden
triangles depicted in Figure 11.1 gives the inflation rules both for the triangles
of TTT and of RD. So far, this does not seem to be too interesting, as one
can do this sort of procedure with the periodic tiling of the plane by squares.
However, in certain cases, such an operation does not result in any loss of
information on the original structure, i.e., it is possible to recover it by an
inverse transformation, also in a local fashion.
Rereading the definitions in the last Sections, one sees that the above
description is precisely what is meant by the following formal definition: a
structure A has an inflation/deflation symmetry related to a similarity trans-
formation T , if T (A) is in the same MLD-class as A (i.e., if T (A) is locally
indistinguishable from a structure which, in turn, is locally equivalent to A; it
is necessary to phrase it this way, because the situation where already T (A)
is locally equivalent to A is too special).
From the last Section, we may conclude that no periodic structure can
have any in-/deflation symmetry related to a nontrivial T , i.e., where T is not
just a rigid motion. The reason can be glimpsed from the following example:
subdividing the square cells of the lattice Z2 into smaller squares of half the
edge length, say, is obviously a local rule. But the converse, re-grouping four
adjacent squares into a bigger square, is not – it requires the knowledge where
the process was started to guarantee fault-free operation, and this means it is
not possible by a local rule. On the other hand, the existence of nontrivial in-
/deflation seems to be a very common feature among the interesting aperiodic
structures [52]. In many cases, T is just a rescaling, but there are important
examples where rotation-dilations are needed, as in certain 2D tilings with
12-fold symmetry (cf. Refs. [55, 5]). We have
Theorem 11.1. The existence of an inflation/deflation symmetry is a prop-
erty of an entire MLD-class: either all members share it, or none has it.
This fact may serve into two directions. Firstly, having established the exis-
tence of in-/deflation for a single structure, one already has in-/deflation for
its entire MLD-class. Secondly, simultaneous (non-)existence of in-/deflation
provides a necessary criterion for two structures to be locally equivalent.
The study of perfect matching rules is another subject where the con-
cept of local equivalence proves fruitful. We say that a structure A possesses
perfect matching rules (essentially in the sense of Ref. [50]) if its LI-class is
determined by the set of its r-patches for some finite radius r (which we
call, if chosen minimally, the matching rule radius of A), i.e., if every other
structure which contains, up to translations, only r-patches which also occur
in A necessarily belongs to the LI-class of A. In the case of LI(PT), a very
simple version in terms of tiles with oriented edges can be given, see [27] for
an illustration. For obvious reasons, this property of a structure is particu-
larly interesting in the case that this structure is supposed to describe the
global order of physically realized structures as quasicrystals [39, 41]. One
may think of a Hamiltonian that favours the patches of the atlas, this way
restricting the groundstate to a member of the LI-class.
It is almost immediate that every structure which is locally equivalent to
one with perfect matching rules must itself possess perfect matching rules:
Theorem 11.2. The existence of perfect matching rules is a property of an
entire MLD-class: either all LI-classes contained in this MLD-class possess
perfect matching rules, or none of the LI-classes can have them.
A Guide to Mathematical Quasicrystals 25
Note, however, that the matching rule radius is not an invariant of MLD-
classes. This is of some relevance in the physical context, if one tries to relate
the matching rules to the local interaction of some suitable Hamiltonian,
i.e. if one searches for a Hamiltonian whose ground states form a specific LI-
class with perfect matching rules. This is due to the fact that the information
contained in a structure may be delocalized (gradually) by the local derivation
of another structure. An estimate for the matching rule radius of a structure
A which is locally equivalent to a structure B with perfect matching rules is
given in [26]. It involves the matching rule radius of B and the relevant radii
for the transition from B to A and vice versa. It is an interesting question
what the infimum of all matching rule radii of the LI-classes inside one MLD-
class is. It has been conjectured that it might actually be zero under some
extra condition. This is rather plausible for systems with inflation-deflation
symmetry, as shrunk-down representatives exist on arbitrarily small scales.
It should be noted that there are certain tilings, such as the Ammann-
Beenker octagonal tiling [1] or Gähler’s dodecagonal “shield” tiling [25], which
do not possess perfect matching rules if one considers only undecorated tiles,
but can be transformed into structures with matching rules by convenient
decorations. In these cases, the introduction of the decoration cannot be
achieved in a local fashion (see Ref. [25]), the naked and decorated tilings
form different MLD-classes and should therefore be distinguished clearly [43].
One might ask for the number of different possibilities to construct tilings
of a given symmetry with perfect matching rules (for a study of the 8-, 10-,
and 12-fold symmetrical cases see Ref. [51]). Here, one is not interested in the
(infinite) variety of representatives of one and the same MLD-class, but in
different MLD-classes with perfect matching rules, such as those defined by
LI(PT) and LI(TTT). At the moment, the only candidate of an infinite family
of different tilings with perfect matching rules is provided by the generalized
Penrose patterns with parameter γ = m + nτ , see [48] for details. A closer
inspection, compare also Ref. [34, 48], shows that these tilings belong to only
one MLD-class, or to two S-MLD-classes – one with fivefold and one with
tenfold symmetry (which contains LI(PT)). This analysis has been extended
in [48] to rational values of γ, which results in an entire tower of LI-classes
that allow a local derivation down the tower, but not upwards. On the bottom
of this tower, we find a well-known friend: the rhombic Penrose tiling, PT.
Similar towers certainly exist for other examples, e.g. in 3-space, but, to our
knowledge, have not yet been analyzed in detail.
along the lines presented above are needed to continue a sound classification
program of aperiodically ordered structures, and for various aspects it is
advantageous not to depend on Fourier transforms.
Let us continue with a speculation. We have seen that there were serious
connections between different examples of patterns with perfect matching
rules. An interesting question is how serious these connections are. From the
past fifteen years of research on quasicrystals and aperiodic order, we have the
feeling that the variety of S-MLD-classes with all magic properties is limited,
if organized properly. To this end, one has to form towers of them, in which
two consecutive members allow a local derivation down the tower, but not up
– as in the case of MLD(TTT) being on top of MLD(PT). With this, we tend
to the following conjecture: the number of towers of quasiperiodic S-MLD-
classes with fixed symmetry, limit translation module of minimal rank13 ,
local inflation/deflation symmetry (with fixed inflation multiplier) and local
perfect matching rules is finite.
This statement is rather fragile: removing essentially any of its conditions,
it is wrong. So, a further exploration of this question (and a proof or disproof)
would be a logical next step in the classification of aperiodic structures –
complementing and perhaps even completing the existing classification of
Fourier modules [36, 53] (mainly based on symmetry alone).
Although a good classification of order, even along these lines, is not in
sight (and it might actually be very far away), one should not close ones
eyes in front of other possibilities. In fact, even if we had the answer to the
above question, it would not be sufficient physically: so far, we have totally,
and deliberately, ignored any stochastic aspect of point sets or tilings. This is
not really tolerable, and the last Section is now devoted to a very brief and
sketchy introduction to a totally different (though, fortunately, not totally
disconnected) universe ...
The first indication on incompleteness of the above approach comes from the
observation that also quasicrystals will show defects (in fact, probably more
than ordinary crystals), and one would like to know the possible scenarios.
As a first step, one can investigate so-called defective vertex configurations
within the geometric setting of a tiling, compare [12] and references therein
for a survey. Although this provides rather interesting insight, and even allows
for some simplistic, but not too unrealistic, models, we shall focus here on an
alternative to perfect tilings or Delone sets that starts from a rather different,
if not antipodal, point of view.
13
Minimal w.r.t. the symmetry, e.g. rank 4 for tenfold symmetry in the plane or
rank 6 for icosahedral symmetry in 3-space.
A Guide to Mathematical Quasicrystals 27
τ +2
s = −νa log(νa ) − νb log(νb ) = log(τ ) ≃ 0.665 . (13.2)
τ +1
This (essentially unrestricted) ensemble has two disadvantages: first, the max-
imal entropy would occur for νa = νb = 1/2 as log(2) ≃ 0.693, and not for
the Fibonacci frequencies. Second, the new ensemble is so random that the
typical ensemble member does no longer show sharp diffraction peaks. Con-
sequently, this model is not suitable to explain Fibonacci-type structures.
To overcome these difficulties, at least in dimensions two and higher,
the idea of a random tiling was put forward by the Cornell group, see [30]
and references therein. Starting from the same proto-tiles as in a perfect
quasiperiodic tiling, one allows all gap-less space-fillings that are face-to-face
and overlap-free, and eventually subject to further restrictions. It is this set
of local, geometric constraints which transforms the unrestricted Bernoulli-
type ensemble into a more interesting and realistic Markov-type ensemble.
E.g., one could start from the two rhombi of the Penrose tiling of the plane
but relax the matching rules. This way, one obtains a random tiling ensemble
with unique entropy maximum at the tile frequencies of the perfect tiling, and
this maximum also corresponds to the unique point (in parameter space) of
maximum symmetry, D10 . So far, the situation seems considerably improved.
However, the diffraction side is not yet fully satisfactory: such planar tiling
ensembles display sharp peaks, but they are, in general, no Bragg peaks.
years after the field took off, it is still in its infancy, and any other statement
bares the risk of delusive security. It is amazing how many things seem to be
known, or almost known, and how few of them are really established, or fully
known. And for many aspects, we have just scratched the surface.
In view of this, I should try to mention some of the points not discussed
above. However, even that would be biased again, and it is probably better
to give the advice to read on in this volume, and to consult the introductory
articles from other summer or winter schools, some of which are given in
the references below. A more complete list of books and proceedings can
be found in a short extra chapter of this book, together with some guiding
comments on what one can find in the titles listed. Anyway, one “take-home”
message should be to stay open-minded to new aspects, to challenge “common
knowledge”, to follow interesting paths, and – last not least – to have fun!
Acknowledgements
References
50. L. S. Levitov, “Local rules for quasicrystals”, Commun. Math. Phys. 119 (1988)
627–66.
51. R. Lück, “Basic ideas of Ammann bar grids”, Int. J. Mod. Phys. B7 (1993)
1437–53.
52. W. F. Lunnon and P. A. B. Pleasants, “Quasicrystallographic tilings”, J.
Math. pures et appl. 66 (1987) 217–63.
53. N. D. Mermin, “(Quasi) crystallography is better in Fourier space”, in: Qua-
sicrystals: The State of the Art, eds. D. P. DiVincenzo and P. J. Steinhardt,
World Scientific, Singapore (1991), 133–83.
54. R. V. Moody (ed.), The Mathematics of Long-Range Aperiodic Order, NATO
ASI Series C 489, Kluwer, Dordrecht (1997).
55. K. Niizeki, “Self-similarity of quasilattices in two dimensions: II. The non-
Bravais-type n-gonal quasilattice”, J. Phys. A22 (1989) 205–18.
56. H.-U. Nissen and C. Beeli, “Electron microscopy of quasicrystals and the va-
lidity of the tiling approach”, Int. J. Mod. Phys. B7 (1993) 1387–413.
57. C.Oguey, M. Duneau and A. Katz, “A geometrical approach of quasiperiodic
tilings”, Commun. Math. Phys. 118 (1988) 99–118.
58. G. van Ophuysen, M. Weber and L. Danzer, “Strictly local growth of Penrose
patterns”, J. Phys. A28 (1995) 281–90.
59. P. A. B. Pleasants, “The construction of quasicrystals with arbitrary symme-
try group”, in: Quasicrystals, eds. C. Janot and R. Mosseri, World Scientific,
Singapore (1995), p. 22–30.
60. R. Penrose, “The rôle of aesthetics in pure and applied mathematical research”,
Bull. Inst. Math. Appl. 10 (1974) 266–71, reprinted in [71].
61. C. Richard, M. Hoeffe, J. Hermisson and M. Baake, “Random tilings: concepts
and examples”, J. Phys. A31 (1998) 6385–408.
62. D. S. Rokhsar, N. D. Mermin and D. C. Wright, “The two-dimensional qua-
sicrystallographic space groups with rotational symmetry less than 23-fold”,
Acta cryst. A44 (1988) 197–211.
63. J. Roth, “The equivalence of two face-centred icosahedral tilings with respect
to local derivability”, J. Phys. A26 (1993) 1455–61.
64. W. Rudin, Functional Analysis, 2nd ed., McGraw Hill, New York (1991).
65. D. Shechtman, I. Blech, D. Gratias and J. Cahn, “Metallic phase with long-
range orientational order and no translational symmetry”, Phys. Rev. Lett. 53
(1984) 1951–3, reprinted in [71].
66. M. Schlottmann, “Geometrische Eigenschaften quasiperiodischer Strukturen”,
Dissertation, Univ. Tübingen (1993); and: “Periodic and quasi-periodic La-
guerre tilings”, Int. J. Mod. Phys. B7 (1993) 1351–63.
67. M. Schlottmann, “Projection formalism for locally compact Abelian groups”,
in: Quasicrystals and Discrete Geometry, ed. J. Patera, Fields Institute Mono-
graphs, vol. 10, AMS, Rhode Island (1998), pp. 247–64.
68. R. L. E. Schwarzenberger, N -dimensional Crystallography, Pitman, San Fran-
cisco (1980).
69. J. E. S. Socolar and P. J. Steinhardt, “Quasicrystals. II. Uni-cell configura-
tions”, Phys. Rev. B34 1986 617–47, reprinted in [71].
70. B. Solomyak, “Dynamics of self-similar tilings”, Ergod. Th. & Dynam. Syst. 17
(1997) 695–738.
71. P. J. Steinhardt and S. Ostlund (eds.), The Physics of Quasicrystals, World
Scientific, Singapore (1987).
72. J. B. Suck, “Prehistory of quasicrystals”, this volume.
73. A. Sütő, “Schrödinger difference equation with deterministic ergodic poten-
tials”, in: Beyond Quasicrystals, eds. F. Axel and D. Gratias, Springer, Berlin
(1995), p. 481–549.