Sei sulla pagina 1di 5

QUASICRYSTAL

A quasiperiodic crystal, or quasicrystal, is a structure that is ordered but not periodic. A


quasicrystalline pattern can continuously fill all available space, but it lacks translational symmetry.
While crystals, according to the classical crystallographic restriction theorem, can possess only
two, three, four, and six-fold rotational symmetries, the Bragg diffraction pattern of quasicrystals
shows sharp peaks with other symmetry orders, for instance five-fold.

Aperiodic tilings were discovered by mathematicians in the early 1960s, and, some twenty years
later, they were found to apply to the study of quasicrystals. The discovery of these aperiodic
forms in nature has produced a paradigm shift in the fields of crystallography. Quasicrystals had
been investigated and observed earlier,[2] but, until the 1980s, they were disregarded in favor of
the prevailing views about the atomic structure of matter. In 2009, after a dedicated search, a
mineralogical finding, icosahedrite, offered evidence for the existence of natural quasicrystals.[3]

Roughly, an ordering is non-periodic if it lacks translational symmetry, which means that a shifted
copy will never match exactly with its original. The more precise mathematical definition is that
there is never translational symmetry in more than n – 1 linearly independent directions, where n
is the dimension of the space filled, e.g., the three-dimensional tiling displayed in a quasicrystal
may have translational symmetry in two dimensions. Symmetrical diffraction patterns result from
the existence of an indefinitely large number of elements with a regular spacing, a property
loosely described as long-range order. Experimentally, the aperiodicity is revealed in the unusual
symmetry of the diffraction pattern, that is, symmetry of orders other than two, three, four, or six.
In 1982 materials scientist Dan Shechtman observed that certain aluminium-manganese alloys
produced the unusual diffractograms which today are seen as revelatory of quasicrystal structures.
Due to fear of the scientific community's reaction, it took him two years to publish the
results[4][5] for which he was awarded the Nobel Prize in Chemistry in 2011.[6]

History

In 1961, Hao Wang asked whether determining if a set of tiles admits a tiling of the plane is an
algorithmically unsolvable problem or not. He conjectured that it is solvable, relying on the
hypothesis that every set of tiles that can tile the plane can do it periodically (hence, it would
suffice to try to tile bigger and bigger patterns until obtaining one that tiles periodically).
Nevertheless, two years later, his student Robert Berger constructed a set of some 20,000 square
tiles (now called Wang tiles) that can tile the plane but not in a periodic fashion. As further
aperiodic sets of tiles were discovered, sets with fewer and fewer shapes were found. In 1976
Roger Penrose discovered a set of just two tiles, now referred to as Penrose tiles, that produced
only non-periodic tilings of the plane. These tilings displayed instances of fivefold symmetry. One
year later Alan Mackay showed experimentally that the diffraction pattern from the Penrose tiling
had a two-dimensional Fourier transform consisting of sharp 'delta' peaks arranged in a fivefold
symmetric pattern.[7] Around the same time Robert Ammann created a set of aperiodic tiles that
produced eightfold symmetry.
Mathematically, quasicrystals have been shown to be derivable from a general method that treats
them as projections of a higher-dimensional lattice. Just as circles, ellipses, and hyperbolic curves
in the plane can be obtained as sections from a three-dimensional double cone, so too various
(aperiodic or periodic) arrangements in two and three dimensions can be obtained from
postulated hyperlattices with four or more dimensions. Icosahedral quasicrystals in three
dimensions were projected from a six-dimensional hypercubic lattice by Peter Kramer and Roberto
Neri in 1984.[8] The tiling is formed by two tiles with rhombohedral shape.

Shechtman first observed ten-fold electron diffraction patterns in 1982, as described in his
notebook.[9] The observation was made during a routine investigation, by electron microscopy, of
a rapidly cooled alloy of aluminium and manganese prepared at the US National Bureau of
Standards (later NIST).

In the summer of the same year Shechtman visited Ilan Blech and related his observation to him.
Blech responded that such diffractions had been seen before.[10][11] Around that time,
Shechtman also related his finding to John Cahn of NIST who did not offer any explanation and
challenged him to solve the observation. Shechtman quoted Cahn as saying: "Danny, this material
is telling us something and I challenge you to find out what it is".

The observation of the ten-fold diffraction pattern lay unexplained for two years until the spring of
1984, when Blech asked Shechtman to show him his results again. A quick study of Shechtman's
results showed that the common explanation for a ten-fold symmetrical diffraction pattern, the
existence of twins, was ruled out by his experiments. Since periodicity and twins were ruled out,
Blech, unaware of the two-dimensional tiling work, was looking for another possibility: a
completely new structure containing cells connected to each other by defined angles and
distances but without translational periodicity. Blech decided to use a computer simulation to
calculate the diffraction intensity from a cluster of such a material without long-range translational
order but still not random. He termed this new structure multiple polyhedral.

The idea of a new structure was the necessary paradigm shift to break the impasse. The “Eureka
moment” came when the computer simulation showed sharp ten-fold diffraction patterns, similar
to the observed ones, emanating from the three-dimensional structure devoid of periodicity. The
multiple polyhedral structure was termed later by many researchers as icosahedral glass but in
effect it embraces any arrangement of polyhedra connected with definite angles and distances
(this general definition includes tiling, for example).

Shechtman accepted Blech's discovery of a new type of material and it gave him the courage to
publish his experimental observation. Shechtman and Blech jointly wrote a paper entitled "The
Microstructure of Rapidly Solidified Al6Mn" [12] and sent it for publication around June 1984 to
the Journal of Applied Physics (JAP). The JAP editor promptly rejected the paper as being better fit
for a metallurgical readership. As a result, the same paper was re-submitted for publication to the
Metallurgical Transactions A, where it was accepted. Although not noted in the body of the
published text, the published paper was slightly revised prior to publication.

Meanwhile, on seeing the draft of the Shechtman–Blech paper in the summer of 1984, John Cahn
suggested that Shechtman's experimental results merit a fast publication in a more appropriate
scientific journal. Shechtman agreed and, in hindsight, called this fast publication "a winning
move”. This paper, published in the Physical Review Letters (PRL),*5+ repeated Shechtman's
observation and used the same illustrations as the original Shechtman–Blech paper in the
Metallurgical Transactions A. The PRL paper, the first to appear in print, caused considerable
excitement in the scientific community.

Next year Ishimasa et al. reported twelvefold symmetry in Ni-Cr particles.[13] Soon, eightfold
diffraction patterns were recorded in V-Ni-Si and Cr-Ni-Si alloys.[14] Over the years, hundreds of
quasicrystals with various compositions and different symmetries have been discovered. The first
quasicrystalline materials were thermodynamically unstable—when heated, they formed regular
crystals. However, in 1987, the first of many stable quasicrystals were discovered, making it
possible to produce large samples for study and opening the door to potential applications. In
2009, following a 10-year systematic search, scientists reported the first natural quasicrystal, a
mineral found in the Khatyrka River in eastern Russia.[3] This natural quasicrystal exhibits high
crystalline quality, equalling the best artificial examples.[15] The natural quasicrystal phase, with a
composition of Al63Cu24Fe13, was named icosahedrite and it was approved by the International
Mineralogical Association in 2010. Furthermore, analysis indicates it may be meteoritic in origin,
possibly delivered from a carbonaceous chondrite asteroid.[16]

Atomic image of a micron-sized grain of the natural Al71Ni24Fe5 quasicrystal (shown in the inset)
from a Khatyrka meteorite. The corresponding diffraction patterns reveal a ten-fold symmetry.[17]

A further study of Khatyrka meteorites revealed micron-sized grains of another natural


quasicrystal, which has a ten-fold symmetry and a chemical formula of Al71Ni24Fe5. This
quasicrystal is stable in a narrow temperature range, from 1120 to 1200 K at ambient pressure,
which suggests that natural quasicrystals are formed by rapid quenching of a meteorite heated
during an impact-induced shock.[17]

Electron diffraction pattern of an icosahedral Ho–Mg–Zn quasicrystal

In 1972 de Wolf and van Aalst[18] reported that the diffraction pattern produced by a crystal of
sodium carbonate cannot be labeled with three indices but needed one more, which implied that
the underlying structure had four dimensions in reciprocal space. Other puzzling cases have been
reported,[19] but until the concept of quasicrystal came to be established, they were explained
away or denied.[20][21] However, at the end of the 1980s the idea became acceptable, and in
1992 the International Union of Crystallography altered its definition of a crystal, broadening it as
a result of Shechtman’s findings, reducing it to the ability to produce a clear-cut diffraction pattern
and acknowledging the possibility of the ordering to be either periodic or aperiodic.[4][notes 1]
Now, the symmetries compatible with translations are defined as "crystallographic", leaving room
for other "non-crystallographic" symmetries. Therefore, aperiodic or quasiperiodic structures can
be divided into two main classes: those with crystallographic point-group symmetry, to which the
incommensurately modulated structures and composite structures belong, and those with non-
crystallographic point-group symmetry, to which quasicrystal structures belong.

Originally, the new form of matter was dubbed "Shechtmanite".[22] The term "quasicrystal" was
first used in print by Steinhardt and Levine[23] shortly after Shechtman's paper was published. The
adjective quasicrystalline had already been in use, but now it came to be applied to any pattern
with unusual symmetry.[notes 2] 'Quasiperiodical' structures were claimed to be observed in
some decorative tilings devised by medieval Islamic architects.[24][25] For example, Girih tiles in a
medieval Islamic mosque in Isfahan, Iran, are arranged in a two-dimensional quasicrystalline
pattern.[26] These claims have, however, been under some debate.[27]

Shechtman was awarded the Nobel Prize in Chemistry in 2011 for his work on quasicrystals. "His
discovery of quasicrystals revealed a new principle for packing of atoms and molecules," stated
the Nobel Committee and pointed that "this led to a paradigm shift within chemistry." [4][28]

Mathematics

A 5-cube as an orthographic projection into 2D using Petrie polygon basis vectors overlaid on the
diffractogram from an icosahedral Ho-Mg-Zn quasicrystal

A 6-cube projected into the rhombic triacontahedron using the golden ratio in the basis vectors.
This is used to understand the aperiodic icosahedral structure of quasicrystals.

There are several ways to mathematically define quasicrystalline patterns. One definition, the "cut
and project" construction, is based on the work of Harald Bohr (mathematician brother of Niels
Bohr). The concept of an almost periodic function (also called a quasiperiodic function) was
studied by Bohr, including work of Bohl and Escanglon.[29] He introduced the notion of a
superspace. Bohr showed that quasiperiodic functions arise as restrictions of high-dimensional
periodic functions to an irrational slice (an intersection with one or more hyperplanes), and
discussed their Fourier point spectrum. These functions are not exactly periodic, but they are
arbitrarily close in some sense, as well as being a projection of an exactly periodic function.

In order that the quasicrystal itself be aperiodic, this slice must avoid any lattice plane of the
higher-dimensional lattice. De Bruijn showed that Penrose tilings can be viewed as two-
dimensional slices of five-dimensional hypercubic structures.[30] Equivalently, the Fourier
transform of such a quasicrystal is nonzero only at a dense set of points spanned by integer
multiples of a finite set of basis vectors (the projections of the primitive reciprocal lattice vectors
of the higher-dimensional lattice).[31] The intuitive considerations obtained from simple model
aperiodic tilings are formally expressed in the concepts of Meyer and Delone sets. The
mathematical counterpart of physical diffraction is the Fourier transform and the qualitative
description of a diffraction picture as 'clear cut' or 'sharp' means that singularities are present in
the Fourier spectrum. There are different methods to construct model quasicrystals. These are the
same methods that produce aperiodic tilings with the additional constraint for the diffractive
property. Thus, for a substitution tiling the eigenvalues of the substitution matrix should be Pisot
numbers. The aperiodic structures obtained by the cut-and-project method are made diffractive
by choosing a suitable orientation for the construction; this is a geometric approach that has also a
great appeal for physicists.

Classical theory of crystals reduces crystals to point lattices where each point is the center of mass
of one of the identical units of the crystal. The structure of crystals can be analyzed by defining an
associated group. Quasicrystals, on the other hand, are composed of more than one type of unit,
so, instead of lattices, quasilattices must be used. Instead of groups, groupoids, the mathematical
generalization of groups in category theory, is the appropriate tool for studying quasicrystals.[32]

Using mathematics for construction and analysis of quasicrystal structures is a difficult task for
most experimentalists. Computer modeling, based on the existing theories of quasicrystals,
however, greatly facilitated this task. Advanced programs have been developed[33] allowing one
to construct, visualize and analyze quasicrystal structures and their diffraction patterns.

Interacting spins were also analyzed in quasicrystals: AKLT Model and 8-vertex model were solved
in quasicrystals analytically.[34]

Study of quasicrystals may shed light on the most basic notions related to quantum critical point
observed in heavy fermion metals. Experimental measurements on the gold-aluminium-ytterbium
quasicrystal have revealed a quantum critical point defining the divergence of the magnetic
susceptibility as temperature tends to zero.[35] It is suggested that the electronic system of some
quasicrystals is located at quantum critical point without tuning, while quasicrystals exhibit the
typical scaling behaviour of their thermodynamic properties and belong to the famous family of
heavy-fermion metals.

Potrebbero piacerti anche