You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/229693397

Ullmann's Encyclopedia of Industrial Chemistry

Chapter · July 2008


DOI: 10.1002/14356007.e22_e01.pub2

CITATIONS READS
13 3,189

3 authors:

Enrique Macia Jean-Marie Dubois


Complutense University of Madrid University of Lorraine
130 PUBLICATIONS 3,201 CITATIONS 357 PUBLICATIONS 7,230 CITATIONS

SEE PROFILE SEE PROFILE

Patricia A Thiel
Iowa State University
412 PUBLICATIONS 14,951 CITATIONS

SEE PROFILE

All content following this page was uploaded by Enrique Macia on 13 August 2020.

The user has requested enhancement of the downloaded file.


Quasicrystals
Standard Article
Enrique Maciá1, Jean-Marie Dubois2, Particia Ann Thiel3
1
Universidad Complutense de Madrid, Madrid, Spain
2
Ecole des Mines, Nancy, France
3
Iowa State University, Ames, Iowa 50011, United State

Copyright # 2008 by Wiley-VCH Verlag GmbH & Co. KGaA. All rights reserved.
DOI: 10.1002/14356007.e22_e01.pub2
Article Online Posting Date: July 15, 2008

Abstract
The article contains sections titled:

1. Introduction
2. Structural Aspects
3. Preparation of Quasicrystalline Phases
4. Chemical Composition
5. Electrical Conductivity
6. Thermal Conductivity and Diffusivity
7. Mechanical and Surface Properties
8. Potential Applications
9. Acknowledgement

1. Introduction
Metals and their alloys have played an important role in the history of human civilization. This is
conveniently illustrated by the fact that two major stages of civilization are known as the Iron Age and
the Bronze Age. This helps to place in context the fundamental importance of the discovery, starting in
1982, of a new family of metallic alloys that triggered a deep revision of our understanding of the
structure of matter [1–6]. This revision ultimately led to the introduction of the concept of order
without periodicity (aperiodic order) and, at the same time, spurred intense interdisciplinary activity
aimed at a deeper understanding of natural systems in which aperiodic order spontaneously arises [7].
To understand the conceptual revolution ignited by the discovery of these alloys, it is necessary to
recall that the traditional classification scheme of solid matter is based on the notion of periodic
ordering of atoms in space. It includes two wide categories: crystalline matter, in which the regular
2 QUASICRYSTALS

location of the constituent atoms determines a long-range order exhibiting translational symmetry;
and amorphous matter, in which, although there are correlations in the positions of neighboring
atoms, long-range order is completely absent. The implicit dichotomy in this classification scheme
made crystalline matter the archetype of order in solid-state physics, while amorphous matter
represented the paradigm of disorder.
This traditional scheme was shaken by the claim that the new alloys, although not conventionally
crystalline, nonetheless exhibited discrete, high-quality diffraction patterns. In other words, they
could not be regarded as amorphous, since their constituent elements showed a remarkable long-range
ordering. The name quasicrystal rapidly became popular for this kind of material [8].
From this discovery, the following picture eventually emerged: In quasicrystals, atoms occupy space
according to a quasiperiodic arrangement instead of the usual periodic fashion observed in crystals. A
simple understanding of what a quasiperiodic arrangement might be can be gained by thinking of the
Fibonacci sequence. There, the nth number is entirely predictable (the sequence is ‘‘ordered’’) but the
rule used to generate it is not analogous to periodicity: each number is the sum of the preceding two,
that is, nj ¼ nj2 þ nj1. (This is in contrast to an arithmetic sequence, in which the rule of generation is
analogous to periodicity.) The geometrical analogue of the Fibonacci sequence constructed by using
two types of segments is thought to represent a ‘‘one-dimensional quasicrystal’’, and real quasicrystals
indeed often exhibit features which can be described successfully in terms of the Fibonacci sequence.
But in addition to their importance in fundamental research, the interest aroused by these new
materials encompassed the field of technological innovations. In fact, the first industrial patent for a
straightforward application of quasicrystals was obtained in 1988 [9]. Since then, the number of
patents related to quasicrystalline phases has progressively increased and, at present, quasicrystal-
containing alloys are produced at a rate of at least several tonnes per annum.
In the years since their discovery, quasicrystalline phases have been observed in many more alloys,
including binary, ternary, and quaternary compounds, mainly based on aluminum and titanium. In
addition to the fivefold symmetry axes characteristic of the icosahedral group, eightfold, tenfold, and
twelvefold symmetries have been discovered, all of which are incompatible with periodic translational
symmetry [3–5]. Figure 1 shows the approximate distribution of the different quasicrystalline alloy
families observed to date, according to their symmetries. The majority belong to the icosahedral group,
followed by the decagonal one.

Figure 1. Percentage distribution of the four major symmetry types of quasicrystalline alloy families
obtained to date
Icosahedral (i) phases exhibit fivefold symmetry, octagonal (o) phases eightfold symmetry, decagonal
(d) phases tenfold symmetry, and dodecagonal (dd) phases twelvefold symmetry
QUASICRYSTALS 3

Due to the increasing number and variety of known quasicrystals, the International Crystal-
lographic Union has redefined the term crystal to include this new kind of order. In this definition, a
crystal is any solid which possesses an essentially discrete diffraction pattern, thus transferring the
essential attribute of crystallinity from real space to reciprocal space. Consequently, within the
crystalline family we can distinguish between periodic crystals and aperiodic crystals. This wider
definition reflects the present idea that microscopic periodicity is a sufficient but not necessary
condition for crystallinity. In this sense, it is interesting to note that the first mention ever made about
using the concept of aperiodic crystals in solid-state physics was probably due to SCHRÖDINGER [10], who
introduced it in his 1944 discussion on the possible nature of genetic material. Finally, an interesting
aspect is the possible relationship between quasicrystalline and conventional crystals, in particular
the possible structural transition from one kind to the other, an aspect that has received a great deal of
attention.

2. Structural Aspects
A feature of quasicrystalline order is the self-similarity of the structure, which constitutes an essential
point in most of the structural models proposed for these materials to date [11–15]. This self-similarity
and, more specifically, the scale on which features appear self-similar, is referred to as the inflation
symmetry.

Self-Similarity. This self-similarity is illustrated in Figure 2, which shows the fivefold arrangement of
spots characteristic of the diffraction pattern of an icosahedral quasicrystal and a geometric con-
struction which bears similar properties. (Under many diffraction conditions, a fivefold symmetry like
that shown in Figure 2 a appears tenfold because of the introduction of a center of inversion symmetry.)
In Figure 2 b, the points of intersection of the lines that join the pentagon vertices define a new
pentagon. Its size has a definite relationship to that of the original one, determined by the ratio t2,
where t is the ratio between the diagonal and the side of the pentagon. This ratio, corresponding to the

Figure 2. a) Low-energy electron diffraction pattern of the fivefold surface of an icosahedral phase of
an Al – Pd – Mn quasicrystal, in which the existence of pentagonal patterns associated with fivefold
axes can be clearly seen [19, 20]. b) Diagram illustrating the self-similarity of the characteristic
diffraction pattern of an icosahedral quasicrystal along a fivefold zone axis.
4 QUASICRYSTALS

irrational number (H5+1)/2, has been known since ancient times as the Golden Mean, and it is
implemented in arts such as painting, sculpture, and architecture. The Golden Mean also describes
fundamental properties of the Fibonacci sequence. It is remarkable that the geometric construction
shown in Figure 2 b, known as a Pitagoric pentagram, is naturally expressed by the spatial atomic
arrangement within a quasicrystal, as can be readily checked by means of a careful comparison with a
real diffraction pattern, such as that shown in Figure 2 a. Another remarkable feature of quasicrystal-
line structures are that they can be ‘‘simplified’’ by lifting them into a higher-dimensional space (higher
than the physical real space in which quasiperiodicity is manifested) [16–18].

Clusters. Clusters containing several tens of atoms are also frequently mentioned in the discussion of
the crystallographic structure of quasicrystals. In particular, Mackay-like clusters and Bergman
clusters have been identified in a number of alloys. It has been proposed that these clusters possess a
special stability, that they constitute the fundamental building blocks of the solid, and that they are
ultimately responsible for all of the unusual physical properties of quasicrystals. This hypothesis
remains under active consideration. Figure 3 a shows the structure of one such cluster which may exist
in icosahedral Al – Pd – Mn, and Fig 3 shows an arrangement of these clusters along the fivefold axis.
Note that the clusters exhibit the same type of inflation symmetry as is illustrated in Figure 2 b.

Shapes. Atomic quasiperiodic order is not just limited to the bulk of the solid, but is present also on its
surface. This is manifest, for instance, in the growth habits of monograins, which in icosahedral

Figure 3. a) Structure of a Mackay icosahedron. Similar clusters can be identified in the bulk of
icosahedral Al – Pd – Mn and Al – Cu – Fe, although other types of clusters can be identified as well. b)
Arrangement of Mackay-type icosahedra showing the hierarchical, self-similar arrangement of
overlapping clusters [12, 14, 15].
QUASICRYSTALS 5

Figure 4. Single grains of quasicrystals resting on a 1  1 mm grid


A) A single grain of Ho – Mg – Zn. The morphology is that of a pentagonal dodecahedron, consistent
with the icosahedral symmetry of the lattice. Photograph courtesy of I. R. Fisher and P. C. Canfield,
Ames Laboratory [26]. B) A single grain of decagonal Al – Ni – Co. The long axis of this sample is
parallel to the tenfold axis, which is perpendicular to the decagonal (aperiodic) planes. The sample has
ten long facetted sides, five of which are visible in the main photograph. The inset shows the face of a
similar crystal cut perpendicular to the tenfold axis, thereby revealing all ten facets. Photograph
courtesy of I. R. Fisher, N. D. Kelso, and P. C. Canfield, Ames Laboratory.

systems often display dodecahedral or triacontahedral shapes [21], as predicted by theory [22, 23]. In
decagonal systems, the growth habits are typically more anisotropic, yet still can reflect the quasi-
crystalline nature. Examples are shown in Figure 4 [24–26] .

3. Preparation of Quasicrystalline Phases


While many quasicrystalline phases are metastable, a number of thermodynamically stable phases
also exist, including Al – Cu – Li, Al – Cu – (Fe, Ru, Os) and Al – Pd – (Mn, Re) [27]. The latter materials
preserve a quasiperiodic structure of extraordinary quality until they melt, which occurs at about
1100 – 1200 K. A detailed knowledge of the phase diagrams of such alloys and the corresponding ability
to precisely determine the boundaries of the small stability regions of the different quasicrystalline
phases have made it possible to grow high-quality single-grain quasicrystals by the conventional
Bridgman and Czochralski growth techniques, and by the less-conventional technique of flux growth
(slow cooling) [27]. There has been considerable progress in the size and quality of samples available.
Optimal single-grain sizes have increased from a few micrometers at their discovery, to about one
centimeter a decade later, to about ten centimeters at present.
The discovery of the thermodynamically stable phases also enabled other conventional techniques
to be used for preparation of quasicrystals in multiple forms, e.g., gas and centrifugal atomization for
preparation of powders; magnetron sputtering for thin films; thermal spray techniques for coatings;
and sintering and pressing for solid bodies, to name but a few.
6 QUASICRYSTALS

4. Chemical Composition
In Figure 5, the various chemical elements found in thermodynamically stable quasicrystalline phases
are shown. The main quasicrystal-forming elements (Al, Ti, Zn, and Cd) are circled.

Figure 5. Chemical elements found in thermodynamically stable quasicrystal alloys. Main forming
elements are circled. The second major constituents are squared. Minor constituents are marked with
a diamond.

Other elements which are major constituents of several quasicrystalline phases include Mg, Ca, Cu,
Zr, Pd, and Yb [27].
The composition of quasicrystalline alloys is quite variable, although it is generally thought to fulfill
particular electron-to-atom ratios, similar to the requirement for Hume – Rothery phases [28–31]. For
the sake of illustration more representative compounds belonging to the icosahedral phase are listed in
the following:
Al-based Zn-based Ti-based Cd-based

Al70Pd20Mn10 Zn80Sc15Mg5 Ti45Zr38Ni17 Cd5.7Yb


Al63Cu25Fe12 Zn60Mg30RE10 Ti40Hf40Ni20 (Ag,In)5.7Ca
Al56Li33Cu11 Zn43Mg37Ga20

Most known stable phases contain three or more elements, but several two-component systems are
notable exceptions: Cd5.7Yb and Cd5.7Ca (both icosahedral) [32] and Ta1.6Te (dodecagonal) [33]. In
QUASICRYSTALS 7

terms of chemical composition, the largest families consist of alloys with about 60 – 75 atom % Al, 40 –
80 atom % Zn, and 40 – 45 atom % Ti.
The compositional range which supports a given thermodynamically stable phase is typically very
small, spanning only a few atom percent. (Because of this, quasicrystals can be appropriately termed
intermetallic compounds.) However, related crystalline phases called approximants usually surround
the quasicrystalline region in the phase diagrams. The approximants bear strong structural resem-
blances to the quasicrystals, and their properties are often similar as well. In terms of properties, the
quasicrystalline phase typically exhibits a local minimum or maximum value, but the adjacent
approximants form a bridge to the ordinary crystalline phases. For some technical applications,
the approximants may be even more appealing than the quasicrystals, since they are less sensitive to
small compositional changes.

5. Electrical Conductivity
Once the existence of a new kind of ordered matter has been established, it seems natural to wonder to
what extent the new structural properties of such materials can determine their transport properties.
In the field of quasicrystals, the most widespread opinion is that quasiperiodic long-range order should
have a significant impact on most physical properties. And, in fact, from the diverse measurements
carried out to date of the electrical resistivity, thermoelectric effect, magnetoresistivity, Hall effect,
specific heat, magnetic susceptibility, and optical properties, it appears that the transport properties of
quasicrystals are, in general, very different from those expected for a simple combination of the
constituent metals [34, 35].
This important result is illustrated in Figure 6, which gives the electrical resistivities r measured at
room temperature for quasicrystals and their constituent (pure) elements. It can be seen that the
isolated elements are good metals possessing low electrical resistivities. In contrast, quasicrystalline
alloys comprised of these elements show electrical resistivities which are higher by several orders of
magnitude, and much higher than typical values for conventional metallic alloys (both crystalline and
amorphous), whose representative values fall between 100 and 300 m Vcm.
In addition, unlike metals, the electrical conductivity s ¼ r1 of quasicrystals increases with
increasing temperature, instead of decreasing, following a power law of the form s / Ta, where
1/2  a  3/2. Also in an anomalous manner, the electrical conductivity decreases when the quasi-
periodic order of the structure is improved, for instance, by means of thermal treatment. And, to
complete the list of oddities, electrical conductivity of quasicrystals depends sensitively on the
chemical composition of the alloy. This has made it possible to observe a metal – insulator transition
for the quaternary icosahedral alloy Al – Pd – Mn – Re, when the Mn content in the alloy is changed by
a mere 5 %.
To summarize, thermodynamically stable quasicrystalline phases are clearly beyond the realm of
metals in terms of their electrical transport properties, and in fact their inclusion in the category of
semiconductor materials has been suggested [36]. In this sense, it is worth mentioning the analogy
between the sensitive dependence of conductivity on chemical composition observed, for instance, for
the icosahedral Al62.5Cu25Fe12.5 alloy, for which an increment of 0.5 % in Al content is enough to double
the conductivity value at low temperature, and the characteristic behavior of conductivity in doped
8 QUASICRYSTALS

Figure 6. Comparison of the electrical resistivities of some stable quasicrystalline phases and their
constituent elements

semiconductors. In this regard, Figure 7 compares the electrical conductivities of several broad classes
of materials.
As with many properties, the effect of quasicrystalline order on transport properties could be
revealed best by comparing properties of samples of similar composition but different structure. The
discovery of stable phases with decagonal symmetry made it possible to perform these measurements
for the same sample, since the decagonal quasicrystals possess periodic order along one direction and
quasiperiodic order within the perpendicular plane. In this way, it was confirmed that the electrical
resistivity in the decagonal alloys Al – Cu – Co and Al – Ni – Co is high and first increases, then
decreases with temperature along the quasiperiodic plane. In contrast, along the periodic axis the
electrical resistivity is lower, and increases with increasing temperature, as is typical for crystalline
metals [37]. This is illustrated in Figure 8 [25]. Thus, the marked anisotropy in the transport
properties must be a consequence of the anisotropy in the sample’s structural order.
Why do aperiodic systems exhibit anomalous electrical transport properties? Realistic models must
encompass two factors. First, is the marked decrease in the number of charge carriers able to
QUASICRYSTALS 9

Figure 7. Comparison of the electrical resistivities of quasicrystals and other broad classes of
materials of technological interest

Figure 8. Anisotropy in the electrical resistivity of decagonal phases [25]


10 QUASICRYSTALS

effectively contribute to the electrical conductivity. This decrease, ultimately due to structural
stabilization, is called a pseudogap in the density of electronic states. Its existence has been confirmed
by means of many experimental techniques [38]. Second is the lack of translational symmetry in these
systems (in two or three dimensions), which precludes the usual understanding of extended electronic
states in a periodic lattice based on Bloch functions [39].
The real importance of the first factor, the pseudogap, is illustrated in Figure 9, which shows
the resistivity r at 300 K as a function of a parameter r, which is defined [40] as the ratio of the
density of states at the Fermi edge to the corresponding free-electron value, that is, r is a direct
measure of the depth of the pseudogap. It turns out that the quasicrystals and their approximants
exhibit a relationship r  r2 at constant temperature, which is obeyed also for disordered systems
[41]. While the pseudogap is certainly important, it is not sufficient to explain all the transport
phenomena, in particular the anomalous temperature dependence of the conductivity noted above.

Figure 9. Resistivity at 300 K versus the ratio of the density of states at the Fermi edge to the
corresponding free-electron value, r. The straight line has a slope of  2. Circles represent icosahedral
quasicrystals (top to bottom: Al – Pd – Re, Al – Pd – Mn, Al – Cu – Ru, and Al – Cu – Fe), and squares
represent approximants in the Al – Cu – Fe system
Courtesy of Jean Marie Dubois, Ecole des Mines, Nancy, France
QUASICRYSTALS 11

6. Thermal Conductivity and Diffusivity


It is well known that most metals are good conductors of heat, and that thermal conductivity is
proportional to the product of their electrical conductivity and temperature (the so-called Wiedemann-
Franz law). None of this holds true for quasicrystals, which are very bad conductors of heat, basically
because the density of free electrons is low. Consequently, heat must propagate by means of atomic
vibrations (phonons). Phonons, which are periodic waves, do not travel easily in an aperiodic lattice
once their wavelength becomes comparable to the size of the constituent atomic clusters. Furthermore,
the vectorial combination of two phonons travels in the opposite direction (an Umklapp process;
German: umklappen ¼ fold back) to that of the incident phonons in quasicrystals, in strong contrast to
the case of regular crystals, in which such a combination is favorable to heat transport. Finally,
quasicrystals contain a large number of atomic sites with two nearly equivalent equilibrium positions.
Such so-called phason sites may be excited by the traveling wave and hence absorb part of its energy.
This also inhibits the conduction of heat.
Some representative values for thermal conductivities (in W m1 K1) of quasicrystals and other
materials of industrial interest are listed in the following:
Ag 420
Cu 380
Al 209
Fe 59
GaAs 40
SiC 35 – 155
TiC 33
TiN 24
Al2O3 27 – 43
Decagonal quasicrystals 1.5 – 2.5
Icosahedral quasicrystals 2
Yttria-doped zirconia 2
Pyrex 1.3

The thermal conductivities of quasicrystals, both decagonal and icosahedral, are clearly lower than
those of thermal insulators that are extensively used in aeronautical industry, such as titanium
carbides or nitrides, doped zirconia, or alumina. In addition, the thermal diffusivity of these new
materials is extraordinarily low, even lower than that of zirconium oxide. And, remarkably, the value
of this diffusivity increases only by a factor of two with increasing temperature until the material
melts.

7. Mechanical and Surface Properties


Table 1 compares mechanical and surface properties of quasicrystals with those of more common
materials. The focus on the quasicrystalline phases in the Al – Pd – Mn and Al – Cu – Fe systems is
12 QUASICRYSTALS

Table 1. Comparison of the properties of quasicrystals with those of common metals and nonmetals

Material Density, Hardness, Fracture Young’s Coefficient of Adhesion energy


g/cm3 HV toughness, modulus, friction a with water,
MPa m1/2 GPa mJ/m2
Quasicrystals
Icosahedral 700 – 900 0.3 200 0.055 72
Al – Pd – Fe
(single grain)
Icosahedral 4.6 800 – 1000 1.6 68 – 85 0.046 55
Al – Cu – Fe
(polygrain)
Common metals
Copper 30 130 105
Aluminum 2.9 25 40 70 0.112 98
High-carbon steel 881 0.071 98
Low-carbon steel 7.8 120 215
Nonmetals
Yttria-doped 5.6
zirconia
Alumina 1950 4.7 300 0.067 120
Diamond >5000 0.04 – 0.05
a
Unlubricated, with respect to a diamond pin in air.

because these data are more abundant than for other quasicrystalline alloys. The quasicrystals are
noteworthy for their hardness, low surface energy, and low friction. An unattractive property is
brittleness, as gauged by their low fracture toughness. Other attractive properties, for which data are
not shown, include oxidation resistance, hydrogen storage capacity, and transition from brittle to
ductile behavior at about three-fourths of the melting point.

Brittleness. Quasicrystals are very brittle. The toughness of single grains is less than 0.5 MPa m1/2. In
contrast, metallic crystals like aluminum are characterized by toughnesses about two orders of
magnitude higher. As a result quasicrystals have little resistance to crack propagation below
450 8C. Hence, they are not very useful as bulk materials but may be more fruitfully employed as
surface coatings or as composites. In such a case, the mechanical integrity is supplied by the substrate
or the matrix, while the quasicrystal offers supplemental functions. For instance, in a coating, this may
be hardness, protection against corrosion, optical absorbance, etc.
However, the brittleness of quasicrystals is overcome at high temperatures. Above about 450 8C,
quasicrystals become plastic and deform more and more with applied stress. Deformation is facilitated
by dislocations, as in conventional crystals. However, in contrast to conventional crystals, there is no
work hardening. This proves the absence of pinning centers and indicates a viscous flow regime that
QUASICRYSTALS 13

might be due to some internal friction between supposedly undeformable atomic entities. Current
research seeks actively to understand the plasticity of quasicrystals. In the same temperature range, it
is also observed that thermal vacancies multiply and play the key role in atomic transport over large
distances. Accordingly, diffusion coefficients of quasicrystals in the temperature range above 450 8C
are similar to those measured in aluminum-based intermetallics. Similarly, the oxidation resistance of
quasicrystals does not significantly depart from that of aluminum alloys or compounds. When in
contact with pure O2 or air, the quasicrystal surface is soon covered by a thin layer of pure alumina.
Again, two temperature regimes may be distinguished, below and above 450 8C, with slow and quick
formation of the oxide film, respectively.
Easy crack-propagation brittleness is indicative of a low cleavage energy (the energy required to
split a crystal into two parts with no plastic deformation; for instance, at very low temperature). This
energy is equal to twice the surface energy of the material, since two fresh surfaces are created by
propagating the crack. In good metals like aluminum, the energy of the clean-surface is typically 1 J/m2
or greater. Very low values in the range of 0.40 – 0.60 J/m2 have been reported recently for Al-based
quasicrystals, assuming a certain form of the analytic dependence of the friction coefficient on surface
energy [42]. Other, more indirect evidence that the surface energy of quasicrystals is indeed small can
be deduced from a comparison of contact angles measured by depositing small liquid droplets on the
surface of a quasicrystalline sample as well as on reference samples such as teflon, metals, or a bulk
oxide such as alumina. By using the approach developed two centuries ago by YOUNG, and making
certain assumptions, the contact angle can be related fairly simply to the reversible adhesion energy
WL of a given liquid onto the surface, larger angles and poorer wetting corresponding to smaller values
of WL.

Water Adhesion. Poor wetting is an appropriate description of the interaction between carefully
polished Al-rich quasicrystals and water. Correspondingly, their reversible adhesion energy with
water is only 25 % larger than that of teflon (55 and 44 mJ/m2, respectively) and one-third of that of
window glass. The fundamental origin of this behavior is not yet understood, but a natural question is
whether it is controlled by the properties of the surface of the quasicrystal, or those of the bulk.
Quasicrystals always carry a thin surface oxide layer, consisting mainly of aluminum oxide [43].
However, bulk alumina is comparable to window glass in that it is wet completely by water, indicating
a high value of WL. Thus, wetting of the quasicrystals is probably not controlled, at least to zeroth order,
by the surface layer, but rather in some way by the underlying atomic and electronic structure of the
bulk. (Note that this discussion is only demonstrably valid for water; common organic solvents wet
quasicrystals well, and this indicates that the wetting may depend on the polarity of the liquid.)

Surface Friction. Surface friction is a further important issue for quasicrystals. As shown in Table 1,
numerous experiments using a spherical diamond pin have demonstrated that friction is lower on a
quasicrystal (m ¼ 0.04 – 0.05) than on steel (m ¼ 0.07) of comparable hardness. With more conventional
riders such as mild steel or WC – Co, the comparison is less favorable which is due to the transfer of
material from the indenter to the surface. During such friction measurements, a major drawback
arises from the brittleness of the contacting bodies and hence formation of wear particles. However,
scratch tracks indicate that quasicrystals subjected to repetitive and severe shear develop some
14 QUASICRYSTALS

ductility during this process, and consequently acquire an ability to self-repair. This brittle-to-ductile
behavior might be associated with the nucleation of crystalline nanometer-sized grains within the
quasicrystalline matrix [44].
The influence of commensurability on friction has been examined by a number of experimental and
theoretical studies [45,46]. In the ideal case, when two workpieces with incommensurate lattices are
brought in contact, the minimal force required to achieve sliding (known as the static frictional force)
vanishes, provided the two substrates are stiff enough [47]. Thus, it has been observed that friction
becomes negligible for incommensurate surfaces sliding under conditions of elastic contact. In real
situations, however, physical contact between two (uncontaminated) surfaces is generally mediated by
third bodies acting like a lubricant film. In that case, the sliding interface should be properly described
in terms of three characteristic lengths, which correspond to the periods of both substrates and the
lubricant layer. Recent theoretical studies indicate that the best low-friction regime is achieved for
incommensurabilities related to cubic irrational numbers rather than to quadratic irrationals, like the
golden mean [48]. According to these results, the low friction observed in quasicrystals cannot be
simply justified in terms of their characteristic Fibonacci-based surface ordering. Quite interestingly,
experimental studies on friction anisotropy of a clean, decagonal AlNiCo quasicrystal, whose surface
terminations exhibit periodic as well as Fibonacci-like atomic ordering along different directions,
reveal a strong connection between interface atomic structure and the mechanisms by which energy is
dissipated [49,50]. This result suggests that electronic and phononic contributions probably play a
significant role in the tribological properties of quasicrystals.

8. Potential Applications
Applications of quasicrystals, both real and anticipated, can be placed into two categories: composites
and coatings. (The term composite is used very broadly to include multiple-phase materials produced
both by artificial mixing and by natural precipitation.) In both forms, the brittleness which char-
acterizes the bulk material is mitigated.
Recent developments for the use of quasicrystals are in the field of catalysis.

8.1. Composites

Steel. Among the composites, a precipitation-strengthened steel is marketed by Sandvik Steel in


Sweden. The precipitates are quasicrystalline. The steel is strong, relatively ductile, corrosion-
resistant, and resistant to overaging. The quasicrystalline precipitates are numerous, and resist
coarsening during aging. Both of these effects must reflect a low interfacial energy. The steel is
attractive for tools in the health industry (surgery, dentistry, acupuncture) and makes up key
components of an electric shaver currently marketed by Philips.

Nanoquasicrystalline Alloys. Another type of bulk composite is an Al-based alloy which can be formed by
rapid solidification and powder processing. The quasicrystals form by precipitation, yielding nanoscale
particles surrounded by an Al matrix. Certain compositions can exhibit especially high strength or
high ductility [51]. One such alloy, Al – Fe – Cr – Ti, exhibits sufficiently high strength at elevated
temperature that it meets an air-force goal level for Al-based alloys. The optimal mechanical properties
QUASICRYSTALS 15

Figure 10. Map of the physical properties of Al quasicrystal composites in comparison with a
conventional Al alloy
Courtesy of Akihisa Inoue, Tohoku University, Japan

of these composites, taken over all icosahedra-forming compositions studied to date, are mapped by the
outer envelope in Figure 10. The properties are compared with those of a more common Al alloy, for
which data are shown by the inner envelope. In all cases (except, of course, thermal expansion), the
quasicrystalline-rich composites excel. Gigas, in Japan, markets these materials as extruded bars and
preformed parts.

Polymers with Quasicrystalline Fillers. A type of composite which is promising, but not yet marketed,
involves polymers with quasicrystalline fillers. Tests with certain high-performance thermoplastic
resins have shown that quasicrystalline AlCuFe particles enhance the wear resistance of the polymers
significantly. This effect is not understood, but may be related to a combination of the hardness, low
friction, and low thermal conductivity intrinsic to the quasicrystal. At the same time, key thermo-
chemical characteristics of the polymer, that is, its glass and melting transitions, are not degraded, and
this indicates that the quasicrystal does not catalyze cross-linking or other disadvantageous reactions
in the resin.

8.2. Coatings
Thermal barriers can be manufactured as thick coatings by thermal spraying techniques or magnetron
sputtering. This application relies upon bulk properties of quasicrystals, namely, low thermal
conductivity and plasticity at high temperature. As such, the quasicrystal is a poor heat conductor
16 QUASICRYSTALS

which is additionally able to resist the shear stress, if any, generated at the interface with the substrate
due to the difference in thermal expansion coefficients. In contrast to zirconia, such a difference is
small, since expansion coefficients for quasicrystals, in the range of (13 – 16)  106 K1, are close to
those of metallic alloys. In comparison to metals, quasicrystals also resist oxidation and corrosion by
sulfur at high temperature. The main drawback to using quasicrystals comes from their rather low
melting temperature and significant atomic mobility. Intercalation of a thin diffusion barrier, typically
made of an oxide such as Y2O3, allows the latter difficulty to be overcome easily. A specific composition
with nearly congruent melting at 1170 8C was developed to enhance the temperature range in which
quasicrystalline thermal barriers may be useful. Although not competitive with doped zirconia above
1050 – 1100 8C, such barriers were successfully tested at 950 8C in a real-time ground test of an aircraft
engine. Application to other combustion engines, to power generators or, in general, to heat insulation
of fast-moving mechanical parts is also of interest.

Coatings for Cookware. The potential application for which quasicrystals have gained most attention is
as coatings which combine wear resistance with low friction and/or low adhesion. A commercial
product which exploited these properties was the quasicrystal-coated frying pan marketed by Sitram.
The quasicrystal coating is not directly competitive with Teflon in terms of its nonstick properties,
since the work of adhesion of the quasicrystal with water, for instance, is not as small as that of Teflon
with water, but adhesion for many foods is reduced relative to a conventional metal pan, and the
coating is much harder than Teflon. The hardness means that normal cutlery can be used with the
quasicrystal coating, an attractive feature relative to Teflon. Furthermore, the low thermal con-
ductivity also presents an advantage in the cookware, since it leads to even surface heating. However,
it appears that technological problems with the manufacturing process remain to be solved, which
points toward the importance of coupling good basic research with technology development in the area
of quasicrystals.
The frying pan is a good example of the general need to balance hardness and surface properties in
coatings. The place of the quasicrystal in the general scheme is illustrated in Figure 11. It can be seen
that the quasicrystal is quite hard, yet offers unique surface properties.
Although one product has been marketed, optimization of coating properties for wear-resistant, low-
friction, and/or low-adhesion applications still requires much technological development. One problem
is that coatings commonly contain significant levels of phase impurities, which can make them more
susceptible to wear by brittle fracture than the pure quasicrystal. The phase impurities arise from the
complex solidification pathway of quasicrystals and from other technical problems of fabrication,
mostly related to the low thermal conductivity. The wear can also lead to a large increase in the friction
coefficient. One possible solution to this problem is the addition of a very small amount of the more
ductile phase Fe – Al to a plasma-sprayed coating of Al – Cu – Fe to improve the coating’s wear
properties. Obviously, the properties of the quasicrystalline coatings are very sensitive to the phase
purity, and more specifically to the exact nature of the secondary phase(s).
QUASICRYSTALS 17

Figure 11. Reversible work of adhesion of water Wwater, derived from contact angle measurements,
as a function of Vickers hardness for various common materials. The size of the symbols shows the
range of properties within each family of materials. The straight line shows the relationship expected
between hardness and reversible work of adhesion: H1/3/W ¼ constant. Such a selection chart empha-
sizes the advantage of quasicrystals in terms of striking a balance between adhesion and hardness.

8.3. Catalysis
Quasicrystalline Al-Cu-Fe is a promising catalyst for steam reforming of methanol [52–55]:

CH3 OH þ H2 O ! CO2 þ 3H2 :

This reaction is an effective way to produce hydrogen under mild reaction conditions, e.g., for fuel
cells. The quasicrystal is prepared as a powder, then crushed to maximize surface area, and finally
leached in NaOH or Na2CO3. The leaching selectively removes Al and Al oxides, leaving each particle
consisting of concentric layers. At its core the solid is quasicrystal, and at its perimeter it consists of
porous Al-OH that supports dispersed nanoparticles of Cu and Fe. The Cu nanoparticles are far more
resistant to sintering than Cu particles in the Cu-based catalysts that are the current industrial
standard.

8.4. Other Applications


Other potential applications not mentioned in this article are hydrogen storage, thermoelectrics
[56–59], and optical absorbers [60]. While promising, these applications seem farther from realization
than those mentioned above.
18 QUASICRYSTALS

Many aspects of the technologies pertaining to quasicrystals have been patented [9, 61–82]. For
more scientific details, see [35, 83–87].
In summary, quasicrystals constitute a fascinating scientific frontier, and a barely-explored source
of new technologies. A central question, which happens to unite the two frontiers, is this: Why does
their remarkable atomic structure lead to unusual physical properties? We believe that the answer to
this question will continue to evolve, hand-in-hand with the evolution of new applications.

9. Acknowledgement
We gratefully acknowledge the financial support which enabled this collaboration. In Madrid, this is
from Universidad Complutense de Madrid through research program PR27/05–14014-BSCH. In
Nancy, this is from the CNRS, local authorities, and PICS No. 545. In Ames, this is from the National
Science Foundation under Grant No. INT-9726785 and through the U.S. Department of Energy under
Contract No. W-405-Eng-82. We also thank M. VICTORIA HERNÁNDEZ and DAVID HERNÁNDEZ for their
generous assistance.

REFERENCES

Specific References
1. D. Shechtman, I. Blech, D. Gratias, J. W. Cahn, Phys. Rev. Lett. 53 (1984) 1951.
2. D. Shechtman, I. Blech, Metall. Trans. A 16 (1985) 1005.
3. T. Ishimas, H.-U. Nissen, Y. Fukano, Phys. Rev. Lett. 55 (1985) 511.
4. L. A. Bendersky, Phys. Rev. Lett. 55 (1985) 1461.
5. N. Wang, H. Chen, K. H. Kuo, Phys. Rev. Lett. 59 (1987) 1010.
6. I. Hargittal, The Chemical Intelligencer 1997, 25 – 48.
7. E. Maciá, Rep. Prog. Phys. 69 (2006) 398 – 441.
8. D. Levine, P. J. Steinhardt, Phys. Rev. Lett. 53 (1984) 2477.
9. FR 8 810 559, 1988 (J. M. Dubois, P. Weinland).
10. E. Schrödinger, What Is Life?, Cambridge University Press, New York 1945.
11. A. Mackay, Nature 344 (1990) 21 – 23.
12. C. Janot, M. de Boissieu, Phys. Rev. Lett. 72 (1994) 1674 – 1677.
13. C. Janot, Phys. Rev. B 53 (1996) 181 – 191.
14. C. Janot, J. Phys.: Condens. Matter 9 (1997) 1493 – 1508.
15. C. Janot, J.-M. Duboisin J.-B. Suck, M. Schreiber, P. Hausler (eds.): An Introduction to Structure,
Physical Properties and Application of Quasicrystalline Alloys, Springer, Berlin 1998.
16. V. Elser, Phys. Rev. B 32 (1985) 4892 – 4898.
17. M. Duneau, A. Katz, Phys. Rev. Lett. 54 (1985) 2688.
18. P. A. Kalugin, A. Y. Kitaev, L. S. Levitov, JETP Lett. 41 (1985) 145.
19. M. Gierer, M. A. Van Hove, A. I. Goldman, Z. Shen, S.-L. Chang, C. J. Jenks, C.-M. Zhang, P. A.
Thiel, Phys. Rev. Lett. 78 (1997) 467 – 470.
20. C. J. Jenks, P. J. Pinhero, Z. Shen, T. A. Lograsso, D. W. Delaney, T. E. Bloomer, S.-L. Chang, C.-
M. Zhang, J. W. Anderegg, A. H. M. Z. Islam, A. I. Goldman, P. A. Thiel in S. Takeuchi, T. Fujiwara
QUASICRYSTALS 19

(eds.): Proceedings of the 6th International Conference on Quasicrystals (ICQ6), World Scientific,
Singapore, 1998, pp. 741 – 748.
21. C. M. H. Beeli, Ph. D. Thesis, Technische Hochschule Zürich (1992).
22. T.-L. Ho in D. P. DiVincenzo, P. Steinhardt (eds.): Quasicrystals: The State of the Art, World
Scientific, Singapore 1991, pp. 403 – 428.
23. J. L. Aragón, F. Dávila, A. Gómez, Phys. Rev. B 51 (1985) 857 – 863.
24. I. R. Fisher, M. J. Kramer, Z. Islam, T. A. Wiener, A. Kracher, A. R. Ross, T. A. Lograsso, A. I.
Goldman, P. C. Canfield, Mat. Sci. Eng. A 10–16 (2000) 294 – 296.
25. I. R. Fisher, M. J. Kramer, Z. Islam, A. R. Ross, A. Kracher, T. Wiener, M. J. Sailer, A. I. Goldman,
P. C. Canfield, Phil. Mag. 79 (1999) 425 – 434.
26. I. R. Fisher, K. O. Cheon, A. F. Panchula, P. C. Canfield, M. Chernikov, H. R. Ott, K. Dennis, Phys.
Rev. B 59 (1999) 308 – 321.
27. A. P. Tsai in Z. M. Stadnik (ed.): Physical Properties of Quasicrystals, Springer, Berlin 1999,
pp. 5 – 50.
28. W. Hume-Rothery, J. Inst. Metals 35 (1926) 295.
29. P. A. Bancel, P. A. Heiney, Phys. Rev. B 33 (1986) 7917.
30. J. Friedel, Helv. Phys. Acta 61 (1988) 538.
31. A. P. Tsai, A. Inoue, T. Masumoto, Mater. Trans. J. Inst. Metals 30 (1990) 463.
32. A. P. Tsai, J. Q. Guo, E. Abe, H. Takahura, T. J. Sato, Nature 408 (2000) 537.
33. M. Conrad, F. Krumeich, B. Harbrecht, Angew. Chem. Int. Ed. 37 (1998) 1384 – 1386.
34. S. Roche, G. Trambly de Laissardiere, D. Mayou, J. Math. Phys. 38 (1997) 1794.
35. Z. M. Stadnik, in M. Cardona, P. Fulde, K. von Klitzing, R. Merlin, H.-J. Queisser, H. Störmer
(eds.): Physical Properties of Quasicrystals, Springer, Berlin 1999.
36. A. Carlsson, Nature 353 (1991) 15 – 6.
37. S. Martin, A. F. Hebard, A. R. Kortan, F. A. Thiel, Phys. Rev. Lett. 67 (1991) 719 – 722.
38. E. Belin-Ferré in Quasicrystals, Materials Research Society Symposium Proceedings vol. 553,
Materials Research Society, Warrendale 1999, p. 347.
39. E. Macia, F. Dominguez-Adame, Electrons, Phonons and Excitons in Low Dimensional Aperiodic
Systems, Editorial Complutense, Madrid 2000.
40. N. F. Mott, Phil. Mag. 19 (1969) 835.
41. U. Mizutani, J. Phys. Condens. Matter 10 (1998) 4609 – 4623.
42. J. M. Dubois, P. Brunet, W. Costin, A. Merstallinger, J. Non-Cryst. Solids 334/335 (2004) 475 –
480.
43. P. A. Thiel, Prog. Surf. Sci. 75 (2004) 69; P. A. Thiel, Prog. Surf. Sci. 75 (2004) 191.
44. J. S. Wu, V. Brien, P. Brunet, C. Dong, J. M. Dubois, Phil. Mag. A 80 (2000) 1645 – 1655.
45. G. He, M. H. Muser, M. O. Robbins, Science 284 (1999) 1650.
46. E. Rajasekaran, X. C. Zeng, D. J. Diestler, in Micro/Nanotribology and Its Applications (B.
Bhushan ed.), Kluwer Academic Publishers, Dordrecht, The Netherlands 1997.
47. O. M. Braun, Y. S. Kivshar, in The Frenkel-Kontorova Model: Concepts. Methods and Applica-
tions, Springer, Berlin (2004); O. M. Braun, Y. S. Kivshar, Phys. Rep. 306 (1998) 1.
48. O. M. Braun, A. Vanossi, E. Tosatti, Phys. Rev. Lett. 95 (2005) 026102.
49. J. Y. Park, D. F. Ogletree, M. Salmeron, R. A. Ribeiro, P. C. Canfield, C. J. Jenks, P. A. Thiel,
Science 309 (2005) 1354.
20 QUASICRYSTALS

50. J. Y. Park, D. F. Ogletree, M. Salmeron, R. A. Ribeiro, P. C. Canfield, C. J. Jenks, P. A. Thiel, Phys.


Rev. B 72 (2005) 220201(R).
51. A. Inoue, H. Kimura, K. Kita in A. I. Goldman, D. J. Sordelet, P. A. Thiel, J. M. Dubois (eds.): New
Horizons in Quasicrystals: Research and Applications, World Scientific, Singapore 1997,
pp. 256 – 263.
52. M. Yoshimura, A. P. Tsai, Journal of Alloys and Compounds 2002, 342, 451.
53. A. P. Tsai, M. Yoshimura, Appl. Cat. A: General 2001, 214, 237.
54. T. Tanabe, S. Kameoka, A. P. Tsai, Catalysis Today 2006, 111, 153.
55. S. Kameoka, T. Tanabe, A. P. Tsai, Catalysis Today 2004, 93-95, 23.
56. A. L. Pope, T. M. Tritt, R. Cagnon, J. Strom-Olsen, Appl. Phys. Lett. 79 (2001) 2345 – 2347.
57. K. Kirihara, K. Kimura, J. Appl. Phys. 92 (2002) 979; T. Nagata, K. Kirihara, K. Kimura, J. Appl.
Phys. 94 (2003) 6560.
58. Y. K. Kuo, J. R. Lai, C. H. Huang, C. S. Lue, T. S. Lin, J. Phys. Condens. Matter 15 (2003) 7555.
59. E. Maciá, Phys. Rev. B 70 (2004) 100201-4(R); E. Maciá, T. Takeuchi, T. Otagiri, Phys. Rev. B 72
(2005) 174208.
60. P. C. Gibbons, K. F. Kelton in Z. M. Stadnik (ed.): Physical Properties of Quasicrystals, Springer,
Berlin 1999, pp. 403 – 432.
61. FR 9 100 549, 1991 (J. M. Dubois, A. Pianelli).
62. FR 9 115 866, 1991 (J. M. Dubois, P. Archaumbault, B. Colleret).
63. FR 9 215 659, 1992 (J. M. Dubois, P. Carthonnet).
64. SE 9 303 280, 1993 (A. H. Stigenberg, J. O. Nilsson, P. Liu).
65. DE 4 425 140, 1994 (T. Eisenhammer, M. Lazarov).
66. FR 9 503 938, 1995 (F. Cyrot-Lackmann, T. Grenet, C. Berger, G. Fourcaudot, C. Gignoux).
67. FR 9 503 939, 1995 (J. M. Dubois, F. Machizaud).
68. JP 24 325 392, 1992 (K. Kita).
69. JP 20 662 692, 1992 (A. Inoue, T. Masumoto, J. Nagahora).
70. JP 6 256 692, 1992 (T. Masumoto, A. Inoue, M. Wattanabe, J. Nagahora, T. Shibata).
71. JP 26 559 193, 1993 (K. Nosaki, T. Masumoto, A. Inoue, T. Yamaguchi).
72. JP 32 437 393, 1993 (T. Sawa, T. Inaba, Y. Takahashi).
73. Samsung Electronics, Ltd., US 5 914 594, 1994 (T. Masumoto, A. Inoue, H. Kimura, Y. Shinohara,
Y. Horio, K. Kita).
74. Sandvik, AB 5 632 826, 1997 (A. Hultin-Stigenberg, J.-O. Nilsson, P. Liu).
75. Iowa State University Research Foundation, US 5 851 317, 1998 (S. B. Biner, D. J. Sordelet, B. K.
Lograsso, I. E. Anderson).
76. Centre National de la Recherche Scientifique, US 5 204 191, 1993 (J.-M. Dubois, P. Weinland).
77. US 5 397 490, 1995 (T. Masumoto, A. Inoue, J. Nagahora).
78. YKK Corp., US 5 419 789, 1995 (K. Kita).
79. Centre National de la Recherche Scientifique, US 5 432 011, 1995 (J.-M. Dubois, A. Pianelli).
80. Iowa State University Research Foundation, US 5 433 978, 1995 (J. E. Shield, A. I. Goldman, I. E.
Anderson, T. W. Ellis, R. W. McCallum, D. J. Sordelet).
81. US 5 458 700, 1995 (T. Masumoto, A. Inoue, M. Watanabe, J. Nagahora, T. Shitata).
82. YKK Corp., US 5 607 523, 1997 (T. Masumoto, A. Inoue, J. Nagahora, T. Shibata, K. Kita).
QUASICRYSTALS 21

83. A. I. Goldman, D. J. Sordelet, P. A. Thiel, J. M. Dubois (eds.): New Horizons in Quasicrystals:


Research and Applications, World Scientific, Singapore 1997.
84. D. J. Sordelet, J. M. Dubois, MRS Bulletin: Quasicrystals , 22 (1997).
85. C. Janot, J.-M. Dubois, Les Quasicristaux. Matie`re à Paradoxes, EDP Sciences, Les Ulis 1998.
86. J.-M. Dubois, P. A. Thiel, A.-P. Tsai, K. Urban (eds.): Quasicrystals, Materials Research Society
Symposium Proceedings vol. 553, Materials Research Society, Warrendale 1999.
87. J. M. Dubois, Useful Quasicrystals, World Scientific, New Jersey 2005.

View publication stats

You might also like