Next Article in Journal
Remote Access Revolution: Chemical Crystallographers Enter a New Era at Diamond Light Source Beamline I19
Next Article in Special Issue
Defects in Static Elasticity of Quasicrystals
Previous Article in Journal
Persistence of Smectic-A Oily Streaks into the Nematic Phase by UV Irradiation of Reactive Mesogens
Previous Article in Special Issue
Methods for Calculating Empires in Quasicrystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structurally Complex Frank–Kasper Phases and Quasicrystal Approximants: Electronic Origin of Stability

by
Valentina F. Degtyareva
* and
Natalia S. Afonikova
Institute of Solid State Physics, Russian Academy of Sciences, Chernogolovka 142432, Russia
*
Author to whom correspondence should be addressed.
Crystals 2017, 7(12), 359; https://doi.org/10.3390/cryst7120359
Submission received: 11 November 2017 / Revised: 26 November 2017 / Accepted: 1 December 2017 / Published: 4 December 2017
(This article belongs to the Special Issue Structure and Properties of Quasicrystalline Materials)

Abstract

:
Metal crystals with tetrahedral packing are known as Frank–Kasper phases, with large unit cells with the number of atoms numbering from hundreds to thousands. The main factors of the formation and stability of these phases are the atomic size ratio and the number of valence electrons per atom. The significance of the electronic energy contribution is analyzed within the Fermi sphere–Brillouin zone interaction model for several typical examples: Cu4Cd3, Mg2Al3 with over a thousand atoms per cell, and for icosahedral quasicrystal approximants with 146–168 atoms per cell. Our analysis shows that to minimize the crystal energy, it is important that the Fermi sphere (FS) is in contact with the Brillouin zones that are related to the strong diffraction peaks: the zones either inscribe the FS or are circumscribed by the FS creating contact at edges or vertices.

1. Introduction

Common metallic structures are based on high-symmetry atomic cells, such as face-centered cubic (fcc), close-packed hexagonal (hcp), and body-centered cubic (bcc), that have a coordination number of either 12 or 8+6. These structures are found in elements in their metallic phases; they are also formed in binary alloys and compounds if constituent elements have small differences in atomic size and electronegativity. A special family of intermetallic alloys is based on tetrahedrally-packed phases, called Frank-Kasper phases, where tetrahedra form polyhedra with coordination numbers of 12 (icosahedron), 14, 15, and 16 [1]. Frank-Kasper polyhedra are basic structural units for many types of metallic alloys, including very complex structures with the number of atoms in the unit cell up to hundreds or thousands, as well as quasicrystals. Structures and properties of complex metallic alloys are considered in several books and review papers [2,3,4,5,6,7].
Crystal structures of very complex intermetallics are built with successive shells of atoms, including polyhedra, like icosahedrons, dodecahedrons, rhombic triacontahedrons, and truncated icosahedrons (soccer ball or “fullercage” [8]). These atomic arrangements are called “Russian doll” (“matryoshka”) clusters [9]. Similar atomic packings exist in icosahedral quasicrystals, known as Mackay-type, Bergman-type, and Tsai-type. These clusters build the structures of quasicrystal approximants with the periodic arrangements in the bcc cell (1/1 type) or more complex periodic structures.
Quasicrystal and approximant phases exist at certain alloy compositions usually defined by valence electron concentrations and represent a kind of Hume-Rothery phases [2,3,4,5]. The crystal energy is lowered by contact of the Fermi level to the Brillouin zone planes formed by strong diffraction peaks. Historically, Hume-Rothery phases were considered first in the Cu-Zn alloys and related binary systems of Cu-group elements with the neighboring elements of higher valences [10,11,12]. Classical Hume-Rothery phases with fccbcc→complex γ-brass→hcp are defined by the number of valence electrons per atom, such as 1.35→1.5→1.62→1.75 [10]. In the case of quasicrystal approximants, these values are usually higher. Diffraction patterns of these compounds consist of several strong diffraction peaks that should be taken into account when considering the stability of their crystal structures.
In the present paper the model of the Fermi-sphere–Brillouin zone (FS–BZ) interaction is applied to complex structures, considering the FS inscribed into the BZ, as well as the FS enveloped inner zones contacting the edges or vertices. Both cases of FS–BZ configuration should affect the band-structure energy and decrease the crystal energy.

2. Theoretical Background and Method of Analysis

Formation of binary compounds at a certain alloy composition is defined by some important factors characterizing the alloy constituents, such as the difference in atomic sizes, electronegativity etc. Beyond these factors, formation of metallic structures is defined by effects of the Fermi sphere–Brillouin zone (FS–BZ) interaction. The Hume-Rothery mechanism has been identified to play a role in the stability of structurally-complex alloy phases, quasicrystals, and their approximants [2,13,14,15,16]. Formation of the complex structures of elemental metals under pressure can also be related to the Hume-Rothery mechanism [17,18,19]. Phase diagrams of binary alloys, such as Au-Cd represent several phases with complex structures that follow the Hume-Rothery rule of stability at certain electron concentrations [20].
The band structure contribution to the crystal structure energy can be estimated by analyzing configurations of Brillouin-Jones zone planes within the nearly free-electron model. A special program, BRIZ, has been developed [21] to construct FS–BZ configurations and to estimate some parameters, such as the Fermi sphere radius (kF), values of reciprocal wave vectors of BZ planes (qhkl), and volumes of BZ and FS. The BZ planes are selected to match the condition qhkl ≈ 2kF, which has a significant structure factor. In this case an energy gap is opened on the BZ plane leading to the lowering of the electron band energy. The ratio of ½qhkl to kF is usually less than 1, and equal to ~0.95; it is called a “truncation” factor. In the FS–BZ presentations by the BRIZ program the BZ planes cross the FS whereas, in the real system, the Fermi sphere is deformed and accommodated inside the BZ due to an increase of the electron density and to a decrease in the electron energy near the BZ plane.
The crystal structure of a phase chosen for the analysis by the BRIZ program is characterized by the lattice parameters and the number of atoms in the unit cell, which define the average atomic volume (Vat). The valence electron concentration (z) is the average number of valence electrons per atom that gives the value of the Fermi sphere radius kF = (3π2z/Vat)1/3. Further structure characterization parameters are the number of BZ planes that are in contact with the FS, the degree of the “truncation” factor, and the value of BZ filling by electronic states, defined as a ratio of the volumes of FS and BZ. Presentations of the FS–BZ configurations are given with the orthogonal axes with the following directions in the common view: a* is looking forward, b* to the right and c* upward.
The FS–BZ interaction for relatively simple structures, such as Cu-Zn phases, is usually described as a contact of the FS to the BZ planes resulting in the decrease of the electronic energy by the formation of the energy gap. In the case of complex structures, diffraction patterns consist of several strong reflections at wave vectors well below 2kF, forming BZs that are inscribed by the FS with contacts at the edges or vertices. These cases are constructed with the BRIZ program by proper choice of presentations: “sphere inside polyhedron” or “polyhedron inside sphere”. Structural stabilization due to the electronic band contribution where FS is enveloping the BZ was discussed for phase stability in In and Sn alloys [22,23,24] and for fcc-based phases in simple metals [25].
Significant arguments for the stability of tetrahedrally-packed structures are related to electrostatic (Ewald or Madelung) contributions to the crystal energy because of the high packing density of atoms in Frank-Kasper polyhedra. These polyhedra are forming interpenetrating building blocks that are packed in the long-range structure with a high symmetry, as, for example, in bcc in the case of 1/1 approximants. The electrostatic energy of ion-ion interactions is usually defined by sums in the real and the reciprocal space providing quick convergence [26,27,28]. In this relation, the configuration of BZs formed by strong reflections and accommodated inside the FS is responsible for the contribution to the Ewald energy and usually prefers high-symmetry polyhedra. The oC16 structure observed in some metals and binary alloys under pressure should be mentioned as an example [29].

3. Results and Discussion

In this work, two groups of complex intermetallic phases are selected: with very large unit cells containing above a thousand atoms, and quasicrystal 1/1 approximants of the Bergman-type and Tsai-type. For our consideration we selected compounds with s and sp valence electron elements (non-transition elements) to allow us to calculate definitely the valence electron concentration. Structural data for compounds are given in Table 1. X-ray diffraction patterns and constructed Brillouin-Jones zones for these structures are presented in Figure 1, Figure 2 and Figure 3. Crystal structure descriptions are given following the Pearson notation [30].

3.1. Giant Unit Cell Compounds Cu4Cd3 and Mg2Al3

Complex compounds with more than 1000 atoms in the unit cell were found in Cu4Cd3 and in Mg2Al3 by Samson [31,32]. Interestingly, the Cu4Cd3 compound differs considerably from isoelectronic phases in neighboring systems of Cu-Zn and Au-Cd that have, in this composition, either bcc or ordered CsCl-type structures [20]. The decisive factor for this difference is that the atomic size ratio does not exceed 10% for Cu/Zn or Au/Cd, but is ∼20% for Cu/Cd. Therefore, in Cu4Cd3, instead of the simple bcc or CsCl structures, a tetrahedrally-packed complex compound is formed.
The diffraction pattern of Cu4Cd3-cF1124 consists of diffraction peaks with relatively strong intensity grouped near the 2kF position (Figure 1a). Below the diffraction pattern, FS–BZ constructions are shown for separate peaks and for the final BZ, including 84 planes in close contact to the FS (lower-right). This consideration confirms the influential role of the Hume-Rothery mechanism for the stability of a complex compound along with an atomic size factor.
A similar approach that considers Brillouin zones in the discussion of tetrahedrally-packed structures is given in [14]. It is necessary to note that, for the Cu4Cd3 compound, the same group of diffraction peaks were selected for the BZ construction (see [14], Figure 40), however, the BZ form is slightly different from the construction made with the BRIZ program more accurate, as given in Figure 1a (lower-right).
The complex compound Mg2Al3 has a similar structure, with 1168 atoms in the unit cell, as was defined by Samson [32] and re-determined recently [33]. Structural data for this phase are given in Table 1 designated as Mg28Al45 following the Pauling file data [40]. The diffraction pattern for this phase (Figure 1b) is similar to that for Cu4Cd3 (Figure 1a) with the same group of strong reflections. However, for Mg28Al45 the position of 2kF for z = 2.62 falls well above the position of strong reflections. This case was discussed by Dubois with the suggestion that, for this compound, “the Fermi sphere overlaps the Jones zone” (see [7], Figure 4). Overlapping Brillouin-Jones zones by the FS for strong reflections are shown in Figure 1b (lower-left, middle). Near the position of 2kF there is a group of reflections forming BZ planes in contact with the inscribed FS (lower-right).

3.2. Quasicrystal Approximants

Interesting groups of tetrahedrally-packed phases represent quasicrystal approximants of the 1/1-type and other types. Some of these phases have been known long before the discovery of quasicrystals: for example, Mg32(Zn,Al)49 and CaCd6 [34,36], and later they were assigned as approximants of icosahedral quasicrystals (QCs). Diffraction patterns of QCs and their approximants have strong reflections at nearly the same positions and both groups of phases exist at very close alloy compositions that can be estimated as valence electron concentrations (z). It is commonly assumed that QCs and their approximants are stabilized by the Hume-Rothery mechanism [2,3,4,5] and exist in regions of z equal to ∼1.8, 2, and ∼2.2–2.3. For those z values, representative QCs-approximants are considered with the structural data listed in Table 1.

3.2.1. Approximants cI168 and cI162

Diffraction patterns for QCs-approximants CaCd6-cI168 and Al30Mg40Zn30-cI162 and FS–BZ configurations are shown in Figure 2. Both phases have similar groups of strong reflections, whereas the 2kF positions are different and are close to the diffraction peak (6 3 1) for z = 2 and to the diffraction peaks (5 4 3), (7 1 0), (5 5 0), for z = 2.3 (Figure 2a,b, respectively).
Strong reflections at lower wave vector values form highly-symmetrical polyhedra that are enveloped by the FS, creating a contact at the edges of the polyhedron with planes (5 0 3) or vertices of the polyhedron with planes (5 3 2), (6 0 0). Due to space group symmetry Im 3 ¯ there is a difference in the intensity for reflections within the same (hkl) set. For example, the structure CaCd6-cI168 results in an intensity ratio of ~6 for reflections (5 0 3) and (5 3 0). Assuming that only the (5 0 3) reflections contribute to the constructions of the BZ, the resulting polyhedron is of the pentagonal dodecahedron type (Figure 2a, lower-left). The BZ polyhedron for (5 3 2) and (6 0 0) reflections represents a rhombic triacontahedron (Figure 2a, lower-middle). Both kinds of polyhedra are inscribed into the FS and with contacts at the vertices. Thus, the polyhedra in reciprocal space are related to polyhedra in real space formed by atom clusters. It should be noted that the BZ that is in contact with the inscribed FS for CaCd6 consists mainly of (6 3 1) and (3 6 1) with nearly the same intensity (Figure 2a, lower-right).
For the approximant Al30Mg40Zn30cI162 the next diffraction peak with indices (5 4 3), (7 1 0), and (5 5 0) forms the BZ planes in essential contact to the FS, as shown in Figure 2b (lower-right). There are 84 facets in the BZ contacting the FS at the same distance. The filling of the BZ by electron states is 0.945 (see Table 1) which satisfies the Hume-Rothery rule well. The same BZ-FS configuration for the Al30Mg40Zn30–approximant was considered in Figure 1 [37] with discussion of matching 2kF to the wave vectors of reflections (5 4 3), (7 1 0), and (5 5 0). In addition to this, it is also necessary to consider the BZs constructed with the reflections (5 0 3) and (5 3 2)/(6 0 0) that are enveloped by the FS, as shown in Figure 2b (lower-left and middle).

3.2.2. Approximants cI160 and cI146

Al30Mg40Zn30-cI162 and Al5CuLi3-cI160 have much similarity in their crystal structures, both belonging to the Bergman-type approximants. Diffraction patterns for both phases are similar, as can be seen from Figure 2b and Figure 3a. The only difference is in the 2kF position: for Al5CuLi3 it is close to the reflection (6 3 1), while for Al30Mg40Zn30 it is close to reflections (5 4 3), (7 1 0), and (5 5 0), as shown in Figure 2b and Figure 3a (lower-right), respectively. It should be noted that, because of space group Im 3 ¯ , the intensity of the (6 1 3) reflection is ~100 times less than that of (6 3 1) and the only (6 3 1) set is participating in the BZ construction. For Al5CuLi3, the BZ filling by electron state is 0.936, matching the Hume-Rothery criteria. In addition to this FS–BZ plane contact, it is also necessary to consider the FS overlap with the inner zones, such as the (5 0 3)-dodecahedron and rhombic triacontahedron formed by (5 3 2) and (6 0 0) planes, which should give a significant contribution to the reduction of the electronic energy.
Our next example of approximants—the phase Au15Cd23Zn11-cI146 was recently found [14] with the valence electron number z = 1.694. For this z value, the 2kF position is close to (5 3 2) and slightly overlaps the (5 0 3) and (6 0 0) reflections, as shown in Figure 3b. The resulting BZ (Figure 3b, lower-right) is completely filled by electron states. It should be noted that BZ construction for this structure was reported (see [14], Figure 34), however, the BRIZ program gives a more accurate construction. Interestingly, for this phase with z ≈ 1.7 there are no other diffraction peaks with comparably high intensity, in contrast to the phases with higher z values.

3.3. Similar Structural Features of Intermetallics with Giant Unit Cells and Approximants

The complex intermetallic phases discussed in the present paper have similar local atomic arrangement based on tetrahedrally-packed Frank-Kasper polyhedra. Relations of atomic configurations organized in different structural types lead to relations in the form of Brillouin-Jones zones. Examples of such similarity provide the BZ constructions for Mg28Al45-cI1168 and Al30Mg40Zn30-cI162 shown in Figure 1b and Figure 2b (lower-right). Diffraction plain sets for BZ construction contacting the FS are as follows:
Mg28Al45(10 8 6)(14 2 0)(10 10 2)
Al30Mg40Zn30(5 4 3)(7 1 0)(5 5 0)
From the comparison of these sets, it can be concluded that the former cell is related to the latter cell as a 2 ×2 × 2 lattice parameter. In this case the number of atoms would be 162 × 8 = 1296, which can be reduced to 1168 by vacancies. It should be noted that, for Mg28Al45, a reflection (10 10 0) expected from this transformation also exists, but has very weak intensity, and for the construction of the BZ, the next reflection (10 10 2) was taken. Formation of a complex supercell by multiplication of lattice parameters was demonstrated by a classical example of γ-brass phase Cu5Zn8. The structure cI52 is formed from bcc with the 2 × 2 × 2 increase of the lattice parameter and an induction of vacancies with some atomic movement that results in an appearance of additional peaks close to the FS, as was discussed by Jones [12]. Interestingly, the forms of the Brillouin zones with faces (8 8 0) and (11 3 3) for Cu4Cd3 and Mg28Al45 are similar (Figure 1) and related to the BZ for γ-brass with faces (3 3 0) and (4 1 1), which indicates the structural relationship between all these compounds. Structural presentations of complex intermetallics as multi-fold supercells of basic metallic close-packed lattices were considered in [25,41].
Structural relations of alloy phases that are built with similar short-range clusters are visible by comparison of BZs related to strong diffraction peaks. For Al5CuLi3-cI160 the outer BZ with faces (5 0 3), (6 0 0), and (5 3 2), shown in Figure 3b (lower-right), represents a polyhedron that looks like the Brillouin-Jones zone of quasicrystals i-AlMgCu and i-AlLiCu constructed from the first strong peaks [6]. It should be noted, that BZs with this type of polyhedron with faces (5 0 3), (6 0 0), and (5 3 2) exist for all approximants discussed where the only difference is in the location either outside or inside the FS. In both cases the FS–BZ interaction is significant for the crystal structure energy.

4. Conclusions

Structurally-complex alloy phases with tetrahedrally-packed polyhedra of Frank-Kasper type are discussed with reference to the free-electron model of Brillouin zone-Fermi sphere interactions. Electron energy reduction originates through the contact of the FS to the BZ planes, as usually assumed for classical Hume-Rothery phases with relatively simple structures. On their diffraction patterns, complex structures display several peaks of strong intensity that produce Brillouin zone polyhedra enveloped by the FS in contact with some edges or vertices. These FS–BZ configurations should be taken into account for the estimation of electron energy as a convincing reason for stability of such complex intermetallic structures. Complex tetrahedrally-packed intermetallic phases consist of successive atomic building blocks arranged as symmetrical polyhedra. The Brillouin-zone constructions for these structures reveal successive BZ polyhedra with nearly the same form as the atomic clusters in the real space. These effects should be regarded by the theoretical considerations.

Acknowledgments

The authors gratefully acknowledge Dr. Olga Degtyareva for valuable discussions and comments. This work is supported by the program “The Matter under High Pressure” of the Russian Academy of Sciences.

Author Contributions

Valentina F. Degtyareva conceived the project and wrote the paper, and Natalia S. Afonikova analyzed the results, and prepared figures and tables.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Frank, F.C.; Kasper, J.S. Complex Alloy structures regarded as sphere packings. I. definitions and basic principles. Acta Cryst. 1958, 11, 184–190. [Google Scholar] [CrossRef]
  2. Mizutani, U. Hume-Rothery Rules for Structurally Complex Alloy Phases; Taylor & Francis: London, UK, 2010. [Google Scholar]
  3. Steurer, W.; Deloudi, S. Crystallography of Quasicrystals-Concepts, Methods and Structures; Springer Series in Materials Science 126; Springer: Berlin, Germany, 2009. [Google Scholar]
  4. Steurer, W.; Deloudi, S. Fascinating quasicrystals. Acta Cryst. A 2008, 64, 1–11. [Google Scholar] [CrossRef] [PubMed]
  5. Tsai, A.P. Discovery of stable icosahedral quasicrystals: progress in understanding structure and properties. Chem. Soc. Rev. 2013, 42, 5352–5365. [Google Scholar] [CrossRef] [PubMed]
  6. Poon, S.J. Electronic properties of quasicrystals an experimental review. Adv. Phys. 1992, 41, 303–363. [Google Scholar] [CrossRef]
  7. Dubois, J.M. Properties and applications of quasicrystals and complex metallic alloys. Chem. Soc. Rev. 2012, 41, 6760–6777. [Google Scholar] [CrossRef] [PubMed]
  8. Dshemuchadse, J.; Steurer, W. More of the “Fullercages”. Z. Anorg. Chem. 2014, 5, 693–700. [Google Scholar] [CrossRef]
  9. Xie, W.; Cava, R.J.; Miller, G.J. Packing of Russian doll clusters to form a nanometer-scale CsCl-type compound in a Cr-Zn-Sn complex metallic alloy. J. Mater. Chem. C 2017, 5, 7215–7221. [Google Scholar] [CrossRef]
  10. Hume-Rothery, W. Researches on the nature, properties, and condition of formation of intermetallic compounds. J. Inst. Met. 1926, 35, 319–335. [Google Scholar]
  11. Mott, N.F.; Jones, H. The Theory of the Properties of Metals and Alloys; Oxford University Press: London, UK, 1936. [Google Scholar]
  12. Jones, H. The Theory of Brillouin Zones and Electron States in Crystals; North Holland Publ.: Amsterdam, The Netherlands, 1962. [Google Scholar]
  13. Berger, R.F.; Walters, P.L.; Lee, S.; Hoffmann, R. Connecting the chemical and physical viewpoints of what determines structure: From 1-D chains to gamma-brasses. Chem. Rev. 2011, 111, 4522–4545. [Google Scholar] [CrossRef] [PubMed]
  14. Lee, S.; Henderson, R.; Kaminsky, C.; Nelson, Z.; Nguyen, J.; Settje, N.F.; Schmid, J.T.; Feng, J. Pseudo-fivefold diffraction symmetries in tetrahedral packing. Chem. Eur. J. 2013, 19, 10244–10270. [Google Scholar] [CrossRef] [PubMed]
  15. Lee, S.; Leighton, C.; Bates, F.S. Sphericity and symmetry breaking in the formation of Frank–Kasper phases from one component materials. Proc. Natl. Acad. Sci. USA 2014, 111, 17723–17731. [Google Scholar] [CrossRef] [PubMed]
  16. Mizutani, U.; Sato, H. The physics of the hume-rothery electron concentration rule. Crystals 2017, 7, 9. [Google Scholar] [CrossRef]
  17. Degtyareva, V.F. Simple metals at high pressures: The Fermi sphere-Brillouin zone interaction model. Phys-Usp 2006, 49, 369–388. [Google Scholar] [CrossRef]
  18. Degtyareva, V.F. Potassium under pressure: Electronic origin of complex structures. Solid State Sci. 2014, 36, 62–72. [Google Scholar] [CrossRef]
  19. Degtyareva, V.F. Structural simplicity and complexity of compressed calcium: Electronic origin. Acta Crystallogr. B 2014, 70, 423–428. [Google Scholar] [CrossRef] [PubMed]
  20. Degtyareva, V.F.; Afonikova, N.S. Complex structures in the Au-Cd alloy system: Hume-Rothery mechanism as origin. Solid State Sci. 2015, 49, 61–67. [Google Scholar] [CrossRef]
  21. Degtyareva, V.F.; Smirnova, I.S. BRIZ: A visualization program for Brillouin zone–Fermi sphere configuration. Z. Kristallogr. 2007, 222, 718–721. [Google Scholar] [CrossRef]
  22. Degtyareva, V.F.; Degtyareva, O.; Winzenick, M.; Holzapfel, W.B. Structural transformations in a simple-hexagonal Hg-Sn alloy under pressure. Phys. Rev. B 1999, 59, 6058–6062. [Google Scholar] [CrossRef]
  23. Degtyareva, V.F.; Winzenick, M.; Holzapfel, W.B. Crystal structure of InBi under pressure up to 75 GPa. Phys. Rev. B 1998, 57, 4975–4978. [Google Scholar] [CrossRef]
  24. Degtyareva, O.; Degtyareva, V.F.; Porsch, F.; Holzapfel, W.B. Face-centered cubic to tetragonal transitions in in alloys under high pressure. J. Phys.-Condens. Mat. 2001, 13, 7295–7303. [Google Scholar] [CrossRef]
  25. Degtyareva, V.F.; Afonikova, N.S. Simple metal and binary alloy phases based on the face centered cubic structure: Electronic origin of distortions, superlattices and vacancies. Crystals 2017, 7, 34. [Google Scholar] [CrossRef]
  26. Smith, A.P.; Ashcroft, N.W. Rapid convergence of lattice sums and structural integrals in ordered and disordered systems. Phys. Rev. B 1988, 38, 12942–12947. [Google Scholar] [CrossRef]
  27. Heine, V.; Weaire, D. Pseudopotential theory of cohesion and structure. Solid State Phys. 1970, 24, 249–463. [Google Scholar]
  28. Weaire, D.; Williams, A.R. On the axial ration of simple hexagonal alloys of tin. Phil. Mag. 1969, 19, 1105–1109. [Google Scholar] [CrossRef]
  29. Degtyareva, V.F. Electronic origin of the orthorhombic Cmca structure in compressed elements and binary alloys. Crystals 2013, 3, 419–430. [Google Scholar] [CrossRef]
  30. Pearson, W.B. The Crystal Chemistry and Physics of Metals and Alloys; Wiley: New York, NY, USA, 1972. [Google Scholar]
  31. Samson, S. The crystal structure of the intermetallic compound Cu4Cd3. Acta Cryst. 1967, 23, 586–600. [Google Scholar] [CrossRef]
  32. Samson, S. The crystal structure of the phase β-Mg2Al3. Acta Cryst. 1965, 19, 401–413. [Google Scholar] [CrossRef]
  33. Feuerbacher, M.; Thomas, C.; Makongo, J.P.A.; Hoffmann, S.; Carrillo-Cabrera, W.; Cardoso, R.; Grin, Y.; Kreiner, G.; Joubert, J.; Schenk, T.; et al. The Samson phase, β-Mg2Al3 revisited. Z. Kristallogr. 2007, 222, 259–288. [Google Scholar]
  34. Bruzzone, G. The Ca-Cd and Ba-Cd systems. Gazz. Chim. Ital. 1972, 102, 234–242. [Google Scholar]
  35. Gomez, C.P.; Lidin, S. Comparative structural study of the disordered MCd6 quasicrystal approximants. Phys. Rev. B 2003, 68, 1–9. [Google Scholar] [CrossRef]
  36. Bergman, G.; Waugh, J.L.; Pauling, L. The crystal structure of the metallic phase Mg32(AI,Zn)49. Acta Cryst. 1957, 10, 254–259. [Google Scholar] [CrossRef]
  37. Sato, H.; Takeuchi, T.; Mizutani, U. Identification of the Brillouin zone planes in the Hume-Rothery matching rule and their role in the formation of the pseudogap from ab initio band calculations for the Al-Mg-Zn 1/1-1/1-1/1 approximant. Phys. Rev. B 2001, 64, 094207. [Google Scholar] [CrossRef]
  38. Guryan, C.A.; Stephens, P.W.; Goldman, A.I.; Gayle, F.W. Structure of icosahedral clusters in cubic Al5.6Li2.9Cu. Phys. Rev. B 1988, 37, 8495–8498. [Google Scholar] [CrossRef]
  39. Audier, M.; Pannetier, J.; Leblanc, M.; Janot, C.; Lang, J.M.; Dubost, B. An approach to the structure of quasicrystals: A single crystal X-ray and neutron diffraction study of the R-Al5CuLi3 phase. Phys. B Condens Mat 1988, 153, 136–142. [Google Scholar] [CrossRef]
  40. Villars, P.; Cenzual, K. (Eds.) Pauling File Binaries Edition; ASM International: Metal Park, OH, USA, 2002. [Google Scholar]
  41. Dshemuchadse, J.; Yung, D.J.; Steurer, W. Structural building principles of complex face-centered cubic intermetallics. Acta Cryst. B 2011, 67, 269–292. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Calculated diffraction patterns for complex intermetallic phases (upper panel) and corresponding Brillouin-Jones zones with the Fermi spheres (lower panel). The position of 2kF for a given valence electron number per atom z and the hkl indices of the planes used for the BZ construction are indicated on the diffraction patterns. (a) Cu4Cd3-cF1124; and (b) Mg28Al45-cF1168. Structural data and FS–BZ evaluation data are given in Table 1.
Figure 1. Calculated diffraction patterns for complex intermetallic phases (upper panel) and corresponding Brillouin-Jones zones with the Fermi spheres (lower panel). The position of 2kF for a given valence electron number per atom z and the hkl indices of the planes used for the BZ construction are indicated on the diffraction patterns. (a) Cu4Cd3-cF1124; and (b) Mg28Al45-cF1168. Structural data and FS–BZ evaluation data are given in Table 1.
Crystals 07 00359 g001
Figure 2. Calculated diffraction patterns for quasicrystal approximants (upper panel) and corresponding Brillouin-Jones zones with the Fermi spheres (lower panel). The position of 2kF for a given valence electron number per atom z and the hkl indices of the planes used for the BZ construction are indicated on the diffraction patterns. (a) CaCd6-cI168; and (b) Al30Mg40Zn30-cI162. Structural data and FS–BZ evaluation data are given in Table 1.
Figure 2. Calculated diffraction patterns for quasicrystal approximants (upper panel) and corresponding Brillouin-Jones zones with the Fermi spheres (lower panel). The position of 2kF for a given valence electron number per atom z and the hkl indices of the planes used for the BZ construction are indicated on the diffraction patterns. (a) CaCd6-cI168; and (b) Al30Mg40Zn30-cI162. Structural data and FS–BZ evaluation data are given in Table 1.
Crystals 07 00359 g002
Figure 3. Calculated diffraction patterns for quasicrystal approximants (upper panel) and corresponding Brillouin-Jones zones with the Fermi spheres (lower panel). The position of 2kF for a given valence electron number per atom z and the hkl indices of the planes used for the BZ construction are indicated on the diffraction patterns. (a) Al5CuLi3-cI160; and (b) Au15Cd23Zn11-cI146. Structural data and FS–BZ evaluation data are given in Table 1.
Figure 3. Calculated diffraction patterns for quasicrystal approximants (upper panel) and corresponding Brillouin-Jones zones with the Fermi spheres (lower panel). The position of 2kF for a given valence electron number per atom z and the hkl indices of the planes used for the BZ construction are indicated on the diffraction patterns. (a) Al5CuLi3-cI160; and (b) Au15Cd23Zn11-cI146. Structural data and FS–BZ evaluation data are given in Table 1.
Crystals 07 00359 g003
Table 1. Structure parameters of several metallic compounds with giant unit cells and quasicrystal approximants. Fermi sphere radius kF, ratios of kF to the distances of the Brillouin zone planes ½qhkl, and the filling degree of the Brillouin zones by electron states VFS /VBZ are calculated by the program BRIZ [21].
Table 1. Structure parameters of several metallic compounds with giant unit cells and quasicrystal approximants. Fermi sphere radius kF, ratios of kF to the distances of the Brillouin zone planes ½qhkl, and the filling degree of the Brillouin zones by electron states VFS /VBZ are calculated by the program BRIZ [21].
PhaseCu4Cd3Mg28Al45CaCd6Al30Mg40Zn30Al5CuLi3Au15Cd23Zn11
Pearson symbolcF1124cF1168cI168cI162cI160cI146
Structural data
Space groupF 4 ¯ 3mFd 3 ¯ mIm 3 ¯ Im 3 ¯ Im 3 ¯ Im 3 ¯
Lattice parameters (Å)a = 25.871a = 28.24a = 15.680a = 14.355a =13.891a = 13.843
Vat.3)15.6719.2822.9518.2616.7518.17
References[31][32,33][34,35][36,37][38,39][14]
FS–BZ data from the BRIZ program
z (number of valence electrons per atom)1.432.6222.32.181.69
kF−1)1.4011.5901.3721.5511.5681.403
Total number
BZ planes
849696849642
hkl: kF/(½qhkl) (880):1.020
(955):1.008
(10.44):1.004
(11.33):0.979
(14.20):1.011
(10.86)
(10.10.2):1.001
(631):1.010
(543):0.968
(701)
(550)
(543):1.002
(701)
(550)
(631):1.022
(543):0.988
(710)
(503):1.060
(600):1.030
(532):1.003
VFS / VBZ0.9550.9660.9500.9450.9361.00

Share and Cite

MDPI and ACS Style

Degtyareva, V.F.; Afonikova, N.S. Structurally Complex Frank–Kasper Phases and Quasicrystal Approximants: Electronic Origin of Stability. Crystals 2017, 7, 359. https://doi.org/10.3390/cryst7120359

AMA Style

Degtyareva VF, Afonikova NS. Structurally Complex Frank–Kasper Phases and Quasicrystal Approximants: Electronic Origin of Stability. Crystals. 2017; 7(12):359. https://doi.org/10.3390/cryst7120359

Chicago/Turabian Style

Degtyareva, Valentina F., and Natalia S. Afonikova. 2017. "Structurally Complex Frank–Kasper Phases and Quasicrystal Approximants: Electronic Origin of Stability" Crystals 7, no. 12: 359. https://doi.org/10.3390/cryst7120359

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop