Skip to content
Publicly Available Published by De Gruyter (O) August 1, 2023

Ni3Sn4 and FeAl2 as vacancy variants of the W-type (“bcc”) structure

  • Andreas Leineweber ORCID logo EMAIL logo

Abstract

Systematization of the vast number of known crystal structures of intermetallic phases is a challenge. One previously proposed group is referred to here as vacancy variants of the W-type structure. Members of this group, may, however, not be easily recognized because of the structural irregularity introduced by the vacancies. Descriptions of the experimentally observed crystal structures of Ni3Sn4 and FeAl2 in terms of vacancy variants of the W-type structure are, respectively, derived by establishing a lattice correspondence with the W-type structure, allowing, in particular, identification of the vacant sites. In both cases only small deviatoric strains are required to obtain the experimentally encountered lattice parameters, and generally small atomic displacements occur from the ideal positions, thus demonstrating significance of the lattice correspondence. The lattice correspondences allow, for both Ni3Sn4 and FeAl2, relating reported microstructure evidence (directions/planes occurring in orientation relationships and crystal habits but also on twinning and slip) with such typical for metals and solid solutions with W-type (“bcc”) structures. This demonstrates that the established lattice correspondences have a significance going beyond a descriptive one, but the underlying W-type structures reveal themselves in the materials’ behavior.

1 Introduction

The huge variety of known crystal structures occurring in inorganic and, in particular, intermetallic solids has always inspired scientists to define groups of structures following certain rules, using geometric, chemical or chemical bonding as criteria. For intermetallic solids see e.g. [1], [2], [3], [4]. One may agree that definition of a group of structures/phases may be regarded as helpful for scientific practice (beyond structure classification) if, on the basis of a small number of principles, as many as possible phases/structures but also some of their properties (beyond pure static structure) can be rationalized and even be predicted. The Hume-Rothery phases are such a group of intermetallic phases showing a series of characteristic crystal structures, the occurrence of which depends, to a large extent, on the electron-to-atom ratio (e/a; often also the term valence electron concentration is used) [5], [6], [7].

One perhaps less prominent group of phases having some overlap with some of the electron-rich Hume-Rothery like phases has been proposed by Schubert [8], [9], [10]. These phases have crystal structures which can be regarded as vacancy variants of the W-type (body centred cubic, bcc) structures. Occurrence of these vacancies has been attributed to the high e/a ratio. The most prominent examples are the γ-brass phases (hence the connection to Hume-Rothery phases) but also Al-rich NiAl. In the case of γ-brass, e.g. the prototypical Cu5Zn8 [11], the 3 × 3 × 3 supercell of the W-type structure has a content of Cu20Zn32Va2, i.e. 1/27 = 3.7 % of the positions are unoccupied (Va, vacancies). In that case, characteristic displacements of the atoms surrounding the vacant positions occur. In case of the CsCl-type (B2) structure of the NiAl phase, on the Al-rich side of the homogeneity range, Ni vacancies (instead of Al atoms substituting Ni) are used to adjust composition [12]. These vacancies can order in Al-rich Ni2Al3 (Ni2Va1Al3) [13], having 1/6 = 16.7 % vacancies. As compared to γ-brass, the displacements around the vacant positions are much smaller. The crystal structures of intermetallic phases identified by Schubert [8], [9], [10] to belong to such vacancy variants of the W-type structure, included fluorite type AuVaGa2 but also CuVaAl2 (25 % vacancies, respectively) but also a series of low-symmetry structures. In particular, such low-symmetry structures may be difficult to identify as W-type related due to the vacancies and the atomic relaxations which may be associated with the “empty” space in the structure.

In the aftermath of Schubert’s works, the idea of a larger group of phases with vacancy variants of the W-type related intermetallic structures has been used occasionally to rationalize crystal structures of quasicrystals and their approximants [3, 14] and of other complicated intermetallic phases [15], [16], [17].

The present work demonstrates that the crystal structures of monoclinic Ni3+xSn4 and of triclinic (ζ-)FeAl2 can be regarded as vacancy variants of the W-type structure. Thereby, it will be shown that microstructure phenomena as orientation-relationships and plastic deformation reported in the literature for these complicated crystal structures shows crystallographic characteristics which can be regarded as inherited from the underlying W-type structure.

2 Structure analysis

We start from a hypothetical W-type structure (index “bcc”) with a conventional unit cell spanned by the basis vectors abcc, bbcc, cbcc, which are formulated with respect to some auxiliary Cartesian coordinate system spanned by the basis vectors e ˆ 1 , e ˆ 2 , e ˆ 3 . One can then write the following matrix equation:

(1) ( a b c c b b c c c b c c ) = ( a b c c 0 0 0 a b c c 0 0 0 a b c c ) ( e ˆ 1 e ˆ 2 e ˆ 3 ) = M b c c T ( e ˆ 1 e ˆ 2 e ˆ 3 ) ,

where the matrix M b c c T contains the components of abcc, bbcc, cbcc in terms of the Cartesian coordinate system as column vectors. This matrix is the inverse transpose of a matrix defined in Ref. [18], hence the choice to name it. abcc is the cubic lattice parameter of that W-type structure.

The basis vectors of some kind of superstructure of the W-type structure, referred to as X may be related with the W-type structure by a lattice correspondence. This lattice correspondence is described by a transformation matrix P, which is used to define a superlattice with basis vectors aX,id, bX,id, cX,id (id for ideal) [19]:[1]

(2) ( a X , i d b X , i d c X , i d ) = ( a b c c b b c c c b c c ) P , M X , i d T = M b c c T P .

P has generally rational components P ij . In order to ensure that aX,id, bX,id, cX,id are translation vectors of the original, body centered W-type structure, 2Pi1, 2Pi2 and 2Pi3 have to be all even or all odd integers, for i = 1, 2, 3.

The lattice parameters aX,id, bX,id, cX,id, αX,id, βX,id, γX,id of such ideal X can straightforwardly be calculated from the components of M X , i d T , but they typically do not agree with the true, e.g. experimentally observed lattice parameters a X , b X , c X , α X , β X , γ X . Such an agreement, however, can be attained by homogeneously straining the lattice of the ideal X by some strain tensor ε:[2]

a X b X c X = I + ε a X , i d b X , i d c X , i d
(3) M X T = ( I + ε ) M X , i d T

where I is the unit matrix. The at maximum 6 independent (or fewer; see example of Ni3Sn4) components of the symmetric strain tensor can be obtained by fitting, requiring agreement of the lattice parameters calculated from M X T and those experimentally observed. As long as no appropriate reference value for the lattice parameter abcc is available, the latter can be taken as fitting parameter and ε is, instead, constrained to be deviatoric/traceless, e.g. by requiring that ε33 = −ε11 − ε22 holds. Thereby, the at maximum 6 components of ε reduce to at maximum only 5.

As e.g. pointed out in works dealing primarily with displacive phase transformations [21, 22], in principle infinitely many lattice correspondences can be constructed relating a parent to a derivative structure, considering the periodicity of the parent structure. Therefore, one may require a maximum agreement of the structures and, in particular, of the translation lattice. The latter agreement can be regarded good, if the only a “small” ε tensor is required to achieve agreement with the true lattice parameters of X. Here we measure this smallness in terms of the Frobenius norm of ε, i.e. square root of the sum of the squares of its components:

(4) x s = ( i , j = 1 3 ε i j 2 ) 1 / 2 .

This value has been suggested by Aizu to represent the extent of a lattice distortion for traceless transformation strain tensors [23], whereby xs can be shown to be independent of the orientation of the tensor ε.

Also, the complete sets of fractional coordinates of the original W-type structure can be included into the transformation. Thereby distances of the atoms remain the same upon transformation by Eq. (2) but the fractional coordinates change due to the basis change. Sometimes a change of origin might have to be considered in order to arrive at a crystallographic description corresponding to one of the standard settings according to Ref. [19]. It is noted that a homogeneous strain due to ε changes the distances, but does not change the fractional coordinates. In any case, the experimental fractional coordinates can be compared with the coordinates resulting from transformation to analyze the distortions of the experimental atomic structure away from the transformed one.

The number of atoms contained in the new unit cell of X is 2 × det(P), i.e. twice the determinant of P, due to the two atoms in the conventional unit cell of the W-type structure. A real number of atoms per unit cell of X lower than 2 × det(P) hints at vacancies of the type mentioned in the introduction. The positions of vacant sites are obtained by comparison of the transformed positions of the atoms of the W-type structure with those experimentally.

When it comes to further discussion of the crystal structure of X, it is helpful to consider prominent lattice directions and planes from the W-type structure. Upon transformation into the new crystallographic coordinate system of X the indices change according to

(5) ( u X v X w X ) = P 1 ( u b c c v b c c w b c c ) .

and

(6) ( h X k X l X ) = ( h b c c k b c c l b c c ) P T .

This can, in particular, be used to interpret crystallographic features in the microstructure of X as inherited from an underlying W-type structure. Note that indices are usually given as row vectors, where we use also notions of the type ( h X / b c c k X / b c c l X / b c c ) = ( h k l ) X / b c c both for directions and planes. Note, further, that the symmetry-equivalent families of directions and planes indicated by indices in ⟨⟩ and {} comprise characteristic combinations of the indices depending on the respective crystal class symmetry.

Many computer programs can handle such transformations. In the present work Vesta [24] was used as tool to do the transformations and also in order to prepare the figures of the atomic structures. This program has also been used to calculate structure factors used in the discussion. The stereographic projections have been drawn by the use of the Crystal Maker suite [25]. Both programs have also been used for intense visual inspection of models of the two crystal structures under consideration. In particular, it was tried to identify directions or planes resembling 〈111〉bcc and/or {110}bcc to eventually elaborate the corresponding P matrices.

3 Ni3Sn4

3.1 Evaluation of crystal structure data from literature

Ni3Sn4 develops readily by the reaction of Ni with Sn somewhat below and also above the melting temperature of Sn, i.e. at temperatures typically relevant during Sn-based soldering [26]. Only at temperatures considerably below the melting point of Sn, different forms of a plate-like intermetallic having with a stoichiometry likely corresponding to NiSn4 develops [26], [27], [28]. While Ni3Sn4 is usually included as one of the major equilibrium phases in phase diagrams Ni–Sn [29], NiSn4 is usually believed to be metastable although final proofs lack for low temperatures.

The crystal structure of Ni3Sn4 has first been determined by Nowotny & Schubert [30, 31], with the crystal structure having monoclinic C2/m symmetry. Jeitschko & Jaberg [32] re-determined the crystal structure using more modern methods and arrived at likely more accurate structure parameters than the previous works. Furuseth & Fjellvåg [33] published single-crystal and powder X-ray diffraction data (SCXRD and PXRD) explicitly dealing with the composition dependence of the structure parameters of Ni3+xSn4 with x = 0 … 0.7. That work in particular presented SXRD data from crystal specimens of compositions of Ni3.08Sn4 and Ni3.39Sn4. Table 1 lists the details for Ni3.08Sn4 and Figure 1(a) illustrates the crystal structure for the purpose of further discussion. It was demonstrated by the structure analyses [33] that it is the filling of the Ni(3) position which is used to arrive at the actual composition Ni3+xSn4. This Ni(3) position is vacant for the ideal Ni3Sn4 composition (Pearson symbol mS14) and is filled up with increasing Ni content. If the position is fully occupied (not attained within the homogeneity range of the monoclinic Ni3Sn4 phase), the composition would be Ni4Sn4 (Pearson symbol mS16). This fully filled up structure is known for CoGe [34] and a polymorph for FeGe [35]. In contrast, in the Ni–Sn system Ni-rich Ni3+xSn4 (NiSn) appears to develop a orthorhombic crystal structure related but distinct from the monoclinic one considered here [3637].

Table 1:

Structure model of Ni3Sn4 (Ni3.08Sn4) due to Furuseth & Fjellvåg [33]. The lattice parameters were reported as a N i 3 S n 4  = 12.199(2) Å, b N i 3 S n 4  = 4.0609(5) Å, c N i 3 S n 4  = 5.2238(7) Å and β N i 3 S n 4  = 105.17(1)° (space group C2/m), and the symmetry-independent regular atomic positions (original labelling) are those listed. Additionally, the transformed fractional coordinates from the W-type structure being inequivalent due to the C2/m symmetry are listed. They are placed and named according to close-by regular positions, except for the position being vacant (Va(1)). d is the distance of the regular positions to the close-by ideal transformed position.

Regular atoms Transformed W-type positions Wyckoff symbol Fractional coordinates x, y, z d(ideal-real) (Å)
Ni(1) 2a 0 0 0
Ni(1,id) 2a 0 0 0 0
Ni(2) 4i 0.21455(6) 0 0.33665(14)
Ni(2,id) 4i 0.2 0 0.3 0.224
Ni(3) 2c 0 0 0.5
Ni(3,id) 2c 0 0 0.5 0
Sn(1) 4i 0.42849(3) 0 0.68636(7)
Sn(1,id) 4i 0.4 0 0.6 0.492
Sn(2) 4i 0.17184(3) 0 0.81235(7)
Sn(2,id) 4i 0.2 0 0.8 0.366

Vacant positions

Va(1) 4i 0.4 0 0.1
Figure 1: 
Structure model of Ni3.08Sn4 as experimentally determined in Ref. [33], compare also Table 1: (a) experimentally determined arrangement of the regular atoms within a unit cell with superposed transformed positions of the underlying W-type structure visible as vertices between “bonds” along the shortest 




1
2



〈
111
〉


b
c
c




$\frac{1}{2}{\langle 111\rangle }_{\mathrm{b}\mathrm{c}\mathrm{c}}$


-like distances. Only the vacant transformed positions (Va(1)) are indicated additionally as small spheres. (b) Same as (a) but without regular atoms, highlighting (red) the rhombic dodecahedron around a single central vertex as typical coordination polyhedron in the W-type structure.
Figure 1:

Structure model of Ni3.08Sn4 as experimentally determined in Ref. [33], compare also Table 1: (a) experimentally determined arrangement of the regular atoms within a unit cell with superposed transformed positions of the underlying W-type structure visible as vertices between “bonds” along the shortest 1 2 111 b c c -like distances. Only the vacant transformed positions (Va(1)) are indicated additionally as small spheres. (b) Same as (a) but without regular atoms, highlighting (red) the rhombic dodecahedron around a single central vertex as typical coordination polyhedron in the W-type structure.

In the previous works the crystal structure of Ni3Sn4 was discussed in relation to those of β-Sn [31, 38] and Ni3Sn2/NiAs [31, 33]. In particular, already in Ref. [31], prominent atomic chains running along the [ 001 ] N i 3 S n 4 directions have been identified in the crystal structure. The idea that these chains might be associated with a 〈111〉bcc direction allowed now, by detailed inspection of the crystal structure, identifying the lattice correspondence between the Ni3Sn4 structure to that of the W-type structure. The basis vectors of an ideal version of Ni3Sn4, a N i 3 S n 4 , i d , b N i 3 S n 4 , i d , and c N i 3 S n 4 , i d spanning its C-centered unit cell can be formulated as (see Eq. (2)):

(7) ( a N i 3 S n 4 , i d b N i 3 S n 4 , i d c N i 3 S n 4 , i d ) = ( a b c c b b c c c b c c ) P = ( a b c c b b c c c b c c ) ( 1 1 1 1 1 1 4 0 1 ) .

The resulting unit cell has a 10-times larger volume than that of the W-type structure, implying 20 ideal positions for atoms per unit cell. For this ideal unit cell it holds

a N i 3 S n 4 , i d = 18 a b c c
b N i 3 S n 4 , i d = 2 a b c c
c N i 3 S n 4 , i d = 3 a b c c
(8) β N i 3 S n 4 , i d = acos 2 3 6 105.79 ° ,

with values already suggesting similarity with the experimental lattice parameters (Table 1), e.g. taking an approximate value of abcc = 3 Å. This agreement is confirmed by the small of entries of the deviatoric strain tensor ε determined (see Eq. (3)):[3]

(9) ε = ε 11 ε 12 ε 13 ε 12 ε 11 ε 13 ε 13 ε 13 2 ε 11 = 0.0050 0.0231 0.0120 0.0231 0.0050 0.0120 0.0120 0.0120 0.0102 .

The value of abcc obtained in the course of the fitting process amounts to 2.9239 Å. The entries of ε yield xs = 0.042 according to Eq. (4). The maximum eigenvalue of ε amounts +0.034 with eigenvector approximately pointing along [ 1 1 0 ] b c c .

Alongside with the unit-cell transformation, rational-valued fractional coordinates of the above-mentioned 20 ideal positions for the atoms are obtained in the course of the cell transformation process. Checks in view of the C2/m symmetry showed that one symmetry-equivalent set of the transformed positions could be taken as origin of the unit cell and that a change in origin was not necessary. Every experimental atomic position can now be associated with one of the ideal transformed positions with distance <0.5 Å, being the inspiration to label the transformed W-type positions. These crystallographically distinct ideal coordinates have been added to Table 1. For one of the (sets of) transformed positions, there are no close-by regular atoms. This position can be regarded as a vacancy which is labelled Va(1); see Figure 1 and Table 1. Actually, also the Ni(3) position is predominantly vacant. If it is fully vacant (composition Ni3Sn4) a fraction of 6/20 = 30 % of the positions of the W-type structure to remain vacant, whereas it is 4/20 = 20 % for the composition Ni4Sn4. Accordingly, Ni3Sn4 can be regarded as a vacancy variant of the W-type structure.

3.2 Observations fitting to the scheme

Table 2 lists the lattice-direction and lattice-plane indices of Ni3Sn4 corresponding to the densest directions and planes of the body-centered cubic W-type structure as obtained from Eqs. (5) and (6). The stereographic projection in Figure 2 depicts a part of these directions and planes for reference purposes. The strong resemblance to a cubic pole figure illustrates the small distortion with respect to the underlying cubic symmetry as implied by the small components of the deviatoric strain tensor in Eq. (9).

Table 2:

Indices for the most prominent (densest) lattice directions and directions of the W-type structure calculated by Eqs. (5) and (6) in terms of the basis of Ni3Sn4 as given by Eq. (7). Corresponding directions and planes are referred in the text as 〈uvwbcc-like and {hkl}bcc-like. For brevity, the indices bcc and Ni3Sn4 have been omitted for the actual entries in the table.

uvwbcc 10 u v w N i 3 S n 4 {hkl}bcc { h k l } N i 3 S n 4
1 2 111 15 1 , 005 , 〈203〉 {110} {020}, {202}, {510}, { 31 2 }
〈100〉 〈154〉, 20 2 {200} {222}, { 80 2 }
〈110〉 〈0 10 0〉, 〈208〉, 15 6 , 352 {211} { 33 2 } , { 11 4 } , { 82 2 } , { 712 } , { 60 4 } , {10 00}, {530}
Figure 2: 
Stereographic projection of the 〈111〉bcc-like directions (black) and {110}bcc-, {100}bcc- and {211}bcc-like plane normals (blue, red and green) of the lattice of Ni3Sn4 (see Table 1) and giving the 





[

u
v
w

]




N
i

3



S
n

4





${[uvw]}_{{\mathrm{N}\mathrm{i}}_{3}{\mathrm{S}\mathrm{n}}_{4}}$


 and 





(

h
k
l

)




N
i

3



S
n

4





${(hkl)}_{{\mathrm{N}\mathrm{i}}_{3}{\mathrm{S}\mathrm{n}}_{4}}$


 using the indices as listed in Table 2 (omitting the parentheses for the plane normals). Open, unindexed symbols present poles on the lower hemisphere. Greater circles pertaining to the 〈111〉bcc-like directions (dotted for lower hemisphere) contain those 





{

1

1
‾

0

}


b
c
c




${\left\{1\overline{1}0\right\}}_{\mathrm{b}\mathrm{c}\mathrm{c}}$


- and 





{

2

1
‾


1
‾


}


b
c
c




${\left\{2\overline{1}\overline{1}\right\}}_{\mathrm{b}\mathrm{c}\mathrm{c}}$


-like planes forming valid slip systems with the corresponding 〈111〉bcc-like direction. Note that the greater circle of 





{
020
}




N
i

3



S
n

4





${\left\{020\right\}}_{{\mathrm{N}\mathrm{i}}_{3}{\mathrm{S}\mathrm{n}}_{4}}$


 (not shown) constitutes a miror plane and the plane normal corresponds to a twofold axis.
Figure 2:

Stereographic projection of the 〈111〉bcc-like directions (black) and {110}bcc-, {100}bcc- and {211}bcc-like plane normals (blue, red and green) of the lattice of Ni3Sn4 (see Table 1) and giving the [ u v w ] N i 3 S n 4 and ( h k l ) N i 3 S n 4 using the indices as listed in Table 2 (omitting the parentheses for the plane normals). Open, unindexed symbols present poles on the lower hemisphere. Greater circles pertaining to the 〈111〉bcc-like directions (dotted for lower hemisphere) contain those { 1 1 0 } b c c - and { 2 1 1 } b c c -like planes forming valid slip systems with the corresponding 〈111〉bcc-like direction. Note that the greater circle of { 020 } N i 3 S n 4 (not shown) constitutes a miror plane and the plane normal corresponds to a twofold axis.

The structure factors for X-rays calculated using the structure model by Furuseth et al. [33] in Table 1 have appreciable magnitudes for {110}bcc- and {200}bcc-like planes, indicating that the underlying W-type structure of the Ni3Sn4 structure strongly influences diffraction pattern of the latter. Higher order and {hkl}bcc-like reflections are, however, considerably weakened by the atomic displacements from the ideal positions. It can also be noted that the {100}bcc- and {111}bcc-like reflections which are expected for a CsCl (B2) superstructure for the W-type, have large structure factors revealing some corresponding atomic ordering of Ni versus Sn. The strain/monoclinic distortion due to Eq. (9) leads to significantly different positions (spitting) for the different representatives of, e.g. the prominent {110}bcc-like reflections in a PXRD pattern.

Yu et al. [39] have observed annealing twins with habit plane { 100 } N i 3 S n 4 having a misorientation corresponding to a 180° rotation around the plane’s normal, a twinning which can also be achieved by shearing along 001 N i 3 S n 4 . According to Table 2, the habit plane is {211}bcc-like whereas the shear direction is 1 2 1 1 1 b c c -like. Indeed, this type of twinning corresponds to the common type of (deformation) twinning in the W-type metals with a shear direction of the parallel to 1 2 1 1 1 b c c [40]. In the same work [39], the same plane and directions of Ni3Sn4 have been identified to form a slip system becoming active upon pillar compression. While the most prominent slip planes for W-type metals are {110}bcc, { 2 11 } b c c is the second most prominent. Both allow for slip along 1 2 1 1 1 b c c [41].

As reported in the course of solidification experiments of Sn droplets on Ni3Sn4 crystal facets [38], primary Ni3Sn4 forms as faceted crystals upon solidification of Sn-rich Ni–Sn melts. These crystals grow as long rods with axis 010 N i 3 S n 4 corresponding to a low-index 〈110〉bcc direction. These rods are faceted on { 100 } N i 3 S n 4 , { 20 1 } N i 3 S n 4 and { 001 } N i 3 S n 4 . While { 100 } N i 3 S n 4 is a {211}bcc-like plane (see above), only higher order {hkl}bcc planes not contained in Table 2 can be identified for { 20 1 } N i 3 S n 4 and { 001 } N i 3 S n 4 so that these cannot easily associated to prominent planes of the W-type structure.

4 FeAl2

4.1 Evaluation of crystal structure data from literature

A first complete atomic structure model of FeAl2 having triclinic P1 symmetry was reported by Corby & Black using SXRD methods [42]. The atomic structure implied 18 atom per unit cell (aP18) with one having mixed occupation by Fe+Al. Bastin et al. [43] investigated the lamellar microstructure resulting from eutectoid decomposition of the high-temperature ε-phase Fe5Al8 consisting into CsCl-type FeAl and FeAl2. Thereby a dual-phase “crystal” specimen was investigated by Weissenberg-camera based SXRD. The authors proposed an A-centered pseudomonoclinic (i.e. triclinic) unit cell for the FeAl2 without referencing to Ref. [42], and without reporting or considering an atomic structure model. Without reference to Bastin et al., but to Corby & Black, Hirata et al. [44] performed transmission electron microscopy (TEM) complemented by selected area electron diffraction (SAED). They also elaborated an orientation relationship (OR) FeAl/FeAl2. A SCXRD-based redetermination of the crystal structure was reported by Chumak et al. [45]. They reported an additional atom per unit cell (aP19) as compared to the model reported by Corby & Black [42]. This additional atom corresponds to one which was already predicted to be present by Mihalkovic & Widom [46] based on DFT calculations, when they analyzed the structure model by Corby & Black. Notably, the structure model offered by Chumak et al. [45] has centrosymmetric triclinic P 1 symmetry.

As pointed out by Scherf et al. [47], the unit cells, which have been reported in Refs. [42, 43, 45], differed but they can be related by basis transformations. In the following we will primarily work with the structure model by Chumak et al. [45], where the basis vectors and other quantities are without primes. The basis vectors of the unit cells formulated by Corby & Black [42] (primed) and Bastin et al. [43] (double primed) can be written in accordance with Ref. [47] as:

(10) ( a F e A l 2 b F e A l 2 c F e A l 2 ) = ( a F e A l 2 b F e A l 2 c F e A l 2 ) ( 1 0 1 0 1 0 0 0 1 ) = ( a F e A l 2 b F e A l 2 c F e A l 2 ) ( 0 1 0 0 0 0.5 1 1 0.5 )

Using in particular EBSD investigations, Scherf et al. [47] reinvestigated the OR FeAl/FeAl2 and showed that it basically agrees with those reported previously [43, 44] if the inversion symmetry of the crystal structure of FeAl2 is taken into account; for more details, see Section 4.2.

In contrast to Ni3Sn4 some relation of the crystal structure of FeAl2 with the W-type structure has been discussed previously in view of close atomic correspondences between certain lattice planes of bcc/CsCl-type FeAl and of FeAl2. This was in particular elaborated by Hirata et al. [44] who used, however, the outdated structure model from Ref. [42]. It was also demonstrated that the electron beam in a TEM can transform FeAl2 particle into a W-type material having the same orientation as the surrounding FeAl matrix [44]. However, an explicit three-dimensional lattice correspondence has not been proposed in these works.

After in-depth visual inspection of the structure model by Chumak et al. [45], a lattice correspondence between the crystal structure of FeAl2 and the W-type structure has now been derived as

(11) ( a F e A l 2 , i d b F e A l 2 , i d c F e A l 2 , i d ) = ( a b c c b b c c c b c c ) P = ( a b c c b b c c c b c c ) ( 1.5 0 0.5 0.5 2 1.5 0.5 1 2.5 ) .

The correspondences for the other sets of basis vectors [42, 43] can be obtained straightforwardly combining Eq. (11) with Eq. (10).

The resulting unit cell due to Eq. (11) has a volume which is 10.5-times larger that of the W-type structure, implying 21 ideal transformed positions for atoms per unit cell. For this ideal (id) unit cell it holds

a F e A l 2 , i d = 2.75 a b c c
b F e A l 2 , i d = 5 a b c c
c F e A l 2 , i d = 8.75 a b c c
α F e A l 2 , i d = acos 1 175 85.66 °
β F e A l 2 , i d = acos 5 77 75.23 °
(12) γ F e A l 2 , i d = acos 1 55 82.25 °

with values already suggesting similarity with the experimental lattice parameters (Table 3), e.g. taking an approximate value of abcc = 3 Å. This agreement is confirmed by the small absolute values of the entries of the deviatoric strain tensor ε determined using the methodology described in Section 2:

(13) ε = ε 11 ε 12 ε 13 ε 12 ε 22 ε 23 ε 13 ε 23 ε 11 ε 22 = 0.0126 0.0015 0.0010 0.0015 0.0173 0.0000 0.0010 0.0000 0.0298 .
Table 3:

Structure model of FeAl2 due to Chumak et al. [45]. The lattice parameters were reported as a F e A l 2  = 4.8745(6) Å, b F e A l 2  = 6.4545(8) Å, c F e A l 2  = 8.7361(10) Å, α F e A l 2  = 87.930(9)°, β F e A l 2  = 74.396(9)°, and γ F e A l 2  = 83.062(9)° (space group P 1 ) and the symmetry-independent regular atomic positions (original labelling) are those listed, named according to close-by regular positions. Additionally, the transformed fractional coordinates from the W-type structure being inequivalent due to the P 1 symmetry are listed. d is the distance of the regular positions to the close-by ideal transformed position.

Regular atoms Transformed W-type positions Wyckoff symbol Fractional coordinates x, y, z d)
Fe(1) 2i 0.1424(2) 0.15941(18) 0.41723(13)
Fe(1,id) 2i 0.14286 0.14286 0.42857 0.140
Fe(2) 2i 0.2297(3) 0.35270(19) 0.87318(14)
Fe(2,id) 2i 0.28571 0.28571 0.85714 0.490
Fe(3) 1a 0 0 0
Fe(3,id) 1a 0 0 0 0
Fe(4)a 2i 0.1632(3) 0.4731(2) 0.59169(17)
Fe(4,id) 2i 0.19048 0.52381 0.57143 0.384
Al(1) 2i 0.4923(6) 0.0078(4) 0.1649(3)
Al(1,id) 2i 0.38095 0.04762 0.14286 0.631
Al(2) 2i 0.0427(6) 0.1135(4) 0.7098(3)
Al(2a,id)b 2i −0.09524 0.23809 0.71429 0.973
Al(2b,id)b 2i 0.23810 −0.09524 0.71429 1.548
Al(3) 2i 0.6015(6) 0.1853(4) 0.5262(3)
Al(3,id) 2i 0.52381 0.19048 0.57143 0.466
Al(4) 2i 0.0189(6) 0.2916(4) 0.1677(3)
Al(4,id) 2i 0.04762 0.38095 0.14286 0.625
Al(5) 2i 0.3095(5) 0.6629(4) 0.0357(3)
Al(5,id) 2i 0.33333 0.66667 0 0.300
Al(6) 2i 0.4221(6) 0.4438(4) 0.2981(3)
Al(6,id) 2i 0.42857 0.42857 0.28571 0.146
  1. aPartially occupied by Al. bSeparate transformed positions, both associated with the regular Al(2) site in the present interpretation due its position somehow intermediate between these transformed positions; see also Figure 3.

The value of abcc obtained in the course of the fitting process amounts 2.9077 Å. The entries of ε yield xs = 0.037 according to Eq. (4). The maximum eigenvalue of ε amounts to +0.030 with eigenvector approximately into [001]bcc direction. The values of xs is quite close to that encountered for Ni3Sn4 (see Section 3.1), but the largest eigenvector is oriented differently.

Figure 3: 
Structure model of FeAl2 as experimentally determined in Ref. [45], compare also Table 3: (a) regular atomic positions within a unit cell with superposed transformed positions of the underlying W-type structure visible as vertices between “bonds” along the shortest 




1
2



〈
111
〉


b
c
c




$\frac{1}{2}{\langle 111\rangle }_{\mathrm{b}\mathrm{c}\mathrm{c}}$


-like distances. Only the positions of Al(2a,id) and Al(2a,id) are additionally indicated additionally as spheres with one single Al(2) atom, resulting on average in a vacancy. (b) Same as (a) but without regular atomic positions, highlighting (red) the rhombic dodecahedron around a single central vertex as typical coordination polyhedron in the W-type structure.
Figure 3:

Structure model of FeAl2 as experimentally determined in Ref. [45], compare also Table 3: (a) regular atomic positions within a unit cell with superposed transformed positions of the underlying W-type structure visible as vertices between “bonds” along the shortest 1 2 111 b c c -like distances. Only the positions of Al(2a,id) and Al(2a,id) are additionally indicated additionally as spheres with one single Al(2) atom, resulting on average in a vacancy. (b) Same as (a) but without regular atomic positions, highlighting (red) the rhombic dodecahedron around a single central vertex as typical coordination polyhedron in the W-type structure.

Alongside with the unit-cell transformation, rational-valued fractional coordinates of the above-mentioned 21 ideal positions for the atoms are obtained by the transformation process. Checks in view of the P 1 symmetry showed that one symmetry-equivalent set of the transformed positions could be taken as origin of the unit cell and that a change in origin was not necessary. Every experimental atomic position, except for that of Al(2), can now be associated with one of the ideal transformed positions with distance lower <0.5 Å, being the inspiration to label the transformed W-type positions. These crystallographically distinct ideal coordinates have been added to Table 3. The regular Al(2) atom assumes a position intermediate between the transformed positions Al(2a,id) (distance 0.97 Å) and Al(2b,id) (1.55 Å); see Table 3. It is these positions which make out the difference of the 19 regular Fe/Al atoms and 21 transformed positions in the unit cells. Accordingly, one obtains a vacancy fraction of 2/21 ≈ 9.5 % in the crystal structure of FeAl2 as reported by Chumak et al. [45]. In view of all this, FeAl2 can be regarded as a vacancy variant of the W-type structure.

4.2 Observations fitting to the scheme

Table 4 lists the lattice-direction and lattice-plane indices of FeAl2 corresponding to the densest directions and planes of the body-centered cubic W-type structure as obtained from Eqs. (5) and (6). The stereographic projection in Figure 4 depicts a part of these directions and planes for reference purposes. The strong resemblance to a cubic pole figure illustrates the small distortion with respect to the underlying cubic symmetry as implied by the small components of the deviatoric strain tensor in Eq. (13). As in the case of Ni3Sn4 (see Section 3.2), also in the case of FeAl2 the X-ray structure factors for the lattice planes corresponding to, in particular {110}bcc are of appreciable magnitude. Thereby, the structure factors for { 221 } F e A l 2 , { 11 3 } F e A l 2 and { 114 } F e A l 2 are significantly larger than for { 1 2 2 } F e A l 2 , { 2 1 2 } F e A l 2 and { 03 1 } F e A l 2 . However, in contrast to Ni3Sn4, {100}bcc- and {111}bcc-like reflections indicative for an underlying B2-type ordering (as in FeAl) are absent.

Table 4:

Indices for the most prominent (densest) lattice directions and directions of the W-type structure calculated by Eqs. (5) and (6) in terms of the basis of FeAl2 as given by Eq. (11). Corresponding directions and planes are referred in the text as 〈uvwbcc-like and {hkl}bcc-like. For brevity, the indices bcc and FeAl2 have been omitted for the actual entries in the table.

uvwbcc 21 u v w F e A l 2 {hkl}bcc { h k l } F e A l 2
1 2 111 813 , 5 2 6 , 7 7 0 , 66 3 {110} {221}, { 1 2 2 } , { 2 1 2 } , { 11 3 } , {114}, { 03 1 }
〈100〉 13 13 , 183 , 2 5 6 {200} { 30 1 } , { 143 } , { 1 2 5 }
〈110〉 14 7 0 , 12 9 6 , 15 6 3 , 2 1 0 , 339 , 1 13 3 {211} {413}, { 3 3 0 } , { 33 2 } , { 2 1 5 } , {335}, {036}, { 250 } , { 1 51 } , { 306 } , {007}, { 2 4 3 } , { 1 4 4 }
Figure 4: 
Stereographic projection of the 〈111〉bcc-like directions (black) and {110}bcc-, {100}bcc- and {211}bcc-like plane normals (blue, red and green) of the lattice of FeAl2 (see Table 3) and giving the 





[

u
v
w

]



F
e
A
l

2




${[uvw]}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


 and 





(

h
k
l

)



F
e
A
l

2




${(hkl)}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


 using the indices as listed in Table 4 (omitting the parentheses for the plane normals). Open, unindexed symbols present poles on the lower hemisphere. Greater circles pertaining to the 〈111〉bcc-like directions (dotted for lower hemisphere) contain those 





{

1

1
‾

0

}


b
c
c




${\left\{1\overline{1}0\right\}}_{\mathrm{b}\mathrm{c}\mathrm{c}}$


- and 





{

2

1
‾


1
‾


}


b
c
c




${\left\{2\overline{1}\overline{1}\right\}}_{\mathrm{b}\mathrm{c}\mathrm{c}}$


-like planes forming valid slip systems with the corresponding 〈111〉bcc-like direction.
Figure 4:

Stereographic projection of the 〈111〉bcc-like directions (black) and {110}bcc-, {100}bcc- and {211}bcc-like plane normals (blue, red and green) of the lattice of FeAl2 (see Table 3) and giving the [ u v w ] F e A l 2 and ( h k l ) F e A l 2 using the indices as listed in Table 4 (omitting the parentheses for the plane normals). Open, unindexed symbols present poles on the lower hemisphere. Greater circles pertaining to the 〈111〉bcc-like directions (dotted for lower hemisphere) contain those { 1 1 0 } b c c - and { 2 1 1 } b c c -like planes forming valid slip systems with the corresponding 〈111〉bcc-like direction.

Scherf et al. [47] showed that the OR FeAl/FeAl2 can be expressed as { 1 01 } F e A l { 114 } F e A l 2 and 111 F e A l 1 1 0 F e A l 2 , and, as mentioned before, that this OR is compatible with previous findings [43, 44]. Moreover, the habit plane was explicitly determined to be { 114 } F e A l 2 , where the authors already emphasized geometric similarity with {110}bcc (as it is the case for the {110}FeAl planes parallel to them according to the OR). Moreover, the parallel directions are the densest packed directions 〈111〉FeAl and 1 1 0 F e A l 2 , the latter one also being 〈111〉bcc-like; see entry 7 7 0 ( F e A l 2 ) (i.e.  1 1 0 F e A l 2 ) in Table 4. As, furthermore, encountered by Scherf et al. [47], the former, experimentally determined OR has also the { 221 } F e A l 2 planes parallel to other {110}FeAl-type planes, whereby already Hirata et al. [44] emphasized that also this plane had structural similarity with {110}FeAl or {110}bcc which was then, furthermore, emphasized by Scherf et al. [47] for { 11 3 } F e A l 2 . It is, in particular, these three (set of) planes of FeAl2 which contain the above-mentioned densely packed 1 1 0 F e A l 2 direction; see Figure 5(a). Using the OR and Table 4, it turns out that all {110}bcc-like planes of FeAl2 are parallel to corresponding planes of FeAl. Accordingly, the OR FeAl/FeAl2 can be regarded (ignoring slight misfits due to the deviatoric strain experienced by the FeAl2) as some kind of cube-on-(pseudo)cube OR. It is, however, emphasized here that the {110}bcc-like planes of FeAl2 not containing the 1 1 0 F e A l 2 direction are not that regular; see Figure 5(b)–(d) versus Figure 5(a). This is well compatible with the larger structure factors of { 221 } F e A l 2 , { 11 3 } F e A l 2 , { 114 } F e A l 2 as compared to { 1 2 2 } F e A l 2 , { 2 1 2 } F e A l 2 , and { 03 1 } F e A l 2 ; see above. In any case, the cube-on-(pseudo)cube OR readily explains the transformation of FeAl2 particles in the electron beam of a TEM into W-type material with the same orientation as the FeAl matrix observed by Hirata et al. [44]. Also the eutectoid microstructure resulting from γ-brass-type Fe5Al8 (itself a vacancy variant of the W-type structure) can now be formulated under the umbrella of a cube-on-cube like OR [47] involving all three phases.

Figure 5: 
Crystal structure of FeAl2 with view along different 〈111〉bcc-type directions and indicated 





{

1

1
‾

0

}


b
c
c




${\left\{1\overline{1}0\right\}}_{\mathrm{b}\mathrm{c}\mathrm{c}}$


-type planes located within the indicated direction (watching downstream that direction): (a) 





[

1

1
‾

0

]



F
e
A
l

2




${[1\overline{1}0]}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


 depicting the 





{
114
}



F
e
A
l

2




${\left\{114\right\}}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


, 





{

11

3
‾


}



F
e
A
l

2




${\left\{11\overline{3}\right\}}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


 and 





{
221
}



F
e
A
l

2




${\left\{221\right\}}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


 planes already discussed as 





{

1

1
‾

0

}


b
c
c




${\left\{1\overline{1}0\right\}}_{\mathrm{b}\mathrm{c}\mathrm{c}}$


-like in Refs. [44, 47], the former two also as habit planes towards FeAl in eutectoid colonies [47, 48], Moreover the 





{
001
}



F
e
A
l

2




${\left\{001\right\}}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


 plane ({211}bcc-like) of the deformation twinning reported in [49] (b) 





[

22

1
‾


]



F
e
A
l

2




${[22\overline{1}]}_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


, (c) 





〈
813
〉



F
e
A
l

2




${\langle 813\rangle }_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


, (d) 





〈

5

2
‾


6
‾


〉



F
e
A
l

2




${\langle 5\overline{2}\overline{6}\rangle }_{{\mathrm{F}\mathrm{e}\mathrm{A}\mathrm{l}}_{2}}$


. Same colors have been used as in Figure 3.
Figure 5:

Crystal structure of FeAl2 with view along different 〈111〉bcc-type directions and indicated { 1 1 0 } b c c -type planes located within the indicated direction (watching downstream that direction): (a) [ 1 1 0 ] F e A l 2 depicting the { 114 } F e A l 2 , { 11 3 } F e A l 2 and { 221 } F e A l 2 planes already discussed as { 1 1 0 } b c c -like in Refs. [44, 47], the former two also as habit planes towards FeAl in eutectoid colonies [47, 48], Moreover the { 001 } F e A l 2 plane ({211}bcc-like) of the deformation twinning reported in [49] (b) [ 22 1 ] F e A l 2 , (c) 813 F e A l 2 , (d) 5 2 6 F e A l 2 . Same colors have been used as in Figure 3.

Strikingly, Li et al. [48] found an OR which can be reformulated for the present purpose as { 1 1 0 } F e A l { 11 3 } F e A l 2 and 111 F e A l 1 1 0 F e A l 2 , with { 11 3 } F e A l 2 identified as habit plane. While the habit plane is different, this OR, nevertheless, implies basically the same cube-on-(pseudo)cube OR as in Ref. [47]. More detailed TEM analysis of FeAl/FeAl2 eutectoid showed supposed antiphase boundaries in FeAl2 as well as misfit dislocations at the phase boundaries, which were unfortunately not characterized with respect to the crystallographic characteristics [49].

Also plastic deformation processes have been characterized for FeAl2 contained in FeAl/FeAl2 eutectoid FeAl/FeAl2 microstructures [48], [49], [50].

Schmitt et al. [49] studied creep at elevated temperature. While FeAl appears to deform first, upon higher strains slip also appears to occur in the FeAl2, where { 221 } 1 1 0 F e A l 2 , { 11 3 } 1 1 0 F e A l 2 , and { 114 } 1 1 0 F e A l 2 were proposed as slip systems. While the structural analogy with { 1 1 0 } 111 F e A l was emphasized based on direct geometrical analysis of the planes and directions (see also above), the described lattice correspondence (Eq. (11)) can be used to show that these slip systems can explicitly be regarded as { 1 1 0 } 111 b c c -like. The authors [49] also directly observed mechanical twins misoriented with respect to the original FeAl2 by 180° rotation along an axis perpendicular to { 001 } F e A l 2 , with this plane also being the habit plane. According to the lattice correspondence this plane is {211}bcc-like (see also Table 4, contained as { 007 } F e A l 2 ), and the twinning observed is an analogue of the typical {211}bcc deformation twins known for W-type metals and solid solutions [40] (also encountered as annealing twins for Ni3Sn4). Moreover, the authors [49] also observed sliding on the {110}bcc-like phase interfaces FeAl/FeAl2.

Li et al. [48, 50] investigated plasticity at room temperature on FeAl/FeAl2 eutectoid microstructures of different interlamellar spacings upon indentation and pillar compression. They identified [50] a plethora of different slip planes by TEM techniques, which include the above-mentioned {110}bcc-like { 221 } F e A l 2 and { 114 } F e A l 2 , but also many others planes which cannot be related with the other {110}bcc- or {211}bcc-like planes contained in Table 4. An evident system behind all these crystallographically distinct slip planes (beyond { 221 } F e A l 2 and { 114 } F e A l 2 ) was neither offered by Li et al. [48], nor could such a system be found by the present author using the insights derivable from the structure relation to the W-type structure. The authors [48] explicitly observed occurrence of sliding on the {110}bcc-like phase interfaces FeAl/FeAl2 upon pillar compression, in particular, also at elevated temperatures [50].

At the end one might return to the pseudomonoclinic character of the crystal structure of FeAl2 proposed by Bastin et al. [43] (double primed basis vectors in Eq. (10)). A closely related unit cell was used in a PXRD work by Balenetskyy et al. [51], with a modified basis with a F e A l 2 = c F e A l 2 , b F e A l 2 = b F e A l 2 and c F e A l 2 = a F e A l 2 . The latter authors emphasized that the crystal structure might truly be monoclinic. Inspection of the atomic structure model due to Chumak et al. [45] published after [51], however, does not reveal hidden symmetry in the atomic structure. Indeed, via the lattice correspondence in Eq. (11) it can be shown that the monoclinic axis corresponds to ± [ 5 59 ] b c c , not containing any symmetry, and that the ideal lattice angles α and γ calculated from the lattice correspondence deviate by more than 1.5° from 90°. Accordingly, to attain truly monoclinic symmetry, new symmetry elements have to build up within the atomic structure, which is, however, not the case in view of the available atomic structure models [42, 45].

5 General discussion

It succeeded to relate the rather low-symmetry crystal structures of monoclinic Ni3Sn4 and triclinic FeAl2, respectively, to an underlying vacancy variant of the W-type crystal structure. The analysis consists of two main steps, which might be valid for every type of derivative structures:

  1. Identification of lattice correspondence (currently by visual inspection of the respective crystal structures) and calculation of transformed coordinates.

  2. Comparison of experimental lattice with the transformed one in terms of deviatoric strain tensor ε and the displacements of the atoms of different types from the ideal positions. This considers the decoration of the individual positions by different atomic species, here including also their non-occupation (vacancies).

Validity of the idea that an experimental structure can be regarded as a derivative structure of a certain (here W-type) parent structure can be demonstrated by smallness of the strain (quantified by xs) and of the atomic displacements.

The xs values calculated from strain tensors for Ni3Sn4 and FeAl2 are smaller than 0.04 (see Sections 3.1 and 4.2). This value can be regarded as small, e.g. in view of a value in the order of somewhat above 0.3 for a volume-independent Bain strain (transformation Cu-type to W-type).[4] Hence, these small xs values support the idea that the structures of Ni3Sn4 and FeAl2 can, indeed, be regarded as vacancy variants of a W-type structure.

Moreover, the atomic displacements from the ideal transformed atomic positions are with <0.5 Å smaller than the average interatomic distances of 3 2 a b c c 2.5 Å . Exception is the Al(2) atoms in FeAl2 which adopt positions in-between two vacant positions. Generally, it is the vacancies within the considered vacancy variants of the W-type structures, which bring a significant level of complexity into the structure and which complicate identification of candidates belonging to such a group of crystal structures.

It may be a task for future work to perform a systematic analysis on details of the crystal structures of vacancy variants of the W-type structure to identify “the system” behind the atomic relaxations around the vacancies and the characteristics of the deviatoric strains (e.g. direction of eigenvectors). However, this requires systematic evaluation of a series of structures, considering the ordering of at least two substitutional elements versus vacancies. The author refrains to perform such an analysis for Ni3Sn4 and FeAl2 alone. Thereby, it has been noted previously by Ilatovskaia et al. [52] that a thermodynamic sublattice model emanating from the structure model by Chumak et al. [45] is apparently unable to produce a thermodynamic description of the FeAl2 phase well reproducing its temperature-dependent homogeneity range. As stated [52], this could indicate that the structure model by Chumak et al. [45] with its mixed-occupied Fe(4) position (see Table 3) might incompletely reflect the composition-dependent crystal structure. The presently elaborated new insight into the characteristics of the crystal structure of FeAl2 might help the extend knowledge about this peculiar phase in the course of future experimental or computational studies on its structure.

Analysis of previously reported experimental observations on Ni3Sn4- and FeAl2-containing materials systems was done under the hypothesis that the same type of lattice directions and lattice planes should be relevant as encountered for W-type metals and solid solutions, and that such directions and lattice planes show up in the corresponding microstructures. Indeed, a significant amount of observations on habit planes of Ni3Sn4 and FeAl2 towards melt or other phases, as well as twinning and slip planes can be reconciled with the hypothesis.

Slip of dislocations in medium complex intermetallic phases constitutes an active area of research, in particular, due to the advent of the potential of micropillar technique allowing observation of slip under crystallographically controlled load conditions even in quite brittle materials [53]. Applying the Peierls−Nabarro model [54], one typically searches for slip systems with short Burgers vectors also in complex crystal structures. The slip systems derived/observed were { 2 1 1 } b c c 111 b c c -like in the case of Ni3Sn4 (see Section 3.2) and { 1 1 0 } b c c 111 b c c -like in the case of FeAl2 (see Section 4.2). The encountered relevant representative of the 〈111〉bcc-like slip direction were 001 N i 3 S n 4 corresponding to  2 × 1 2 111 b c c for Ni3Sn4 and 1 1 0 F e A l 2 corresponding to 3 × 1 2 111 b c c for FeAl2 in view of Table 2 and Table 4. These are, indeed, the, respectively, shortest translation periodicities along 〈111〉bcc-like directions in Ni3Sn4 and FeAl2.

Hence, the evident significance of crystallographic characteristics known for W-type metals and solid solutions in Ni3Sn4 and FeAl2 for the materials’ behavior going beyond the pure crystal structure geometry justifies the initially geometrically derived idea to regard the crystal structures of these phases to the larger group vacancy variants of W-type related crystal structures as introduced in Schubert’s works [8], [9], [10].

6 Conclusions

The previously reported low-symmetry crystal structures of Ni3Sn4 and FeAl2 were shown to constitute members of the crystal structure group of vacancy variants of the W-type structure. The analysis involved finding a lattice correspondence and ideal positions of the atoms as well as of vacant positions. It is shown that only small deviatoric strains are necessary to obtain the experimentally observed lattice parameters, and that the atomic displacements from the ideal positions are generally small, except occasionally close to the vacant positions.

The obtained lattice correspondences between the lattices of Ni3Sn4 and FeAl2 and the W-type structure allow reinterpretation of previous microstructure evidence concerning orientation relationships and habit planes with other phases, as well as concerning twinning and slip systems. This showed that the observed lattice directions and planes correspond to low-index/densely packed directions and planes of the underlying W-type/“bcc” structure.


Corresponding author: Andreas Leineweber, Institute of Materials Science, TU Bergakademie Freiberg, Gustav-Zeuner-Str. 5, 09599 Freiberg, Germany, E-mail:

  1. Author contributions: The author has accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: None declared.

  3. Conflict of interest statement: The author declares no conflicts of interest regarding this article.

References

1. Schubert, K. Kristallstrukturen Zweikomponentiger Phasen; Springer: Berlin, Heidelberg, 1964.10.1007/978-3-642-94904-3Search in Google Scholar

2. Kripyakevich, P. I. Structure Types of Intermetallic Compounds; Nauka: Moscow, 1977.Search in Google Scholar

3. Steurer, W., Dshemuchadse, J. Intermetallics: structures, properties, and statistics. In Oxford Science Publications, 1st ed.; Oxford University Press: Oxford, Vol. 26, 2016.10.1093/acprof:oso/9780198714552.001.0001Search in Google Scholar

4. Pöttgen, R., Johrendt, D. Intermetallics: Synthesis, Structure, Function; De Gruyter: Berlin, Boston, 2019.10.1515/9783110636727Search in Google Scholar

5. Massalski, T. B., King, H. W. Alloy phases of the noble metals. Prog. Mater. Sci. 1963, 10, 3–78. https://doi.org/10.1016/0079-6425(63)90008-2.Search in Google Scholar

6. Massalski, T., Mizutani, U. Electronic structure of Hume-Rothery phases. Prog. Mater. Sci. 1978, 22, 151–262. https://doi.org/10.1016/0079-6425(78)90001-4.Search in Google Scholar

7. Mizutani, U. Introduction to the Electron Theory of Metals; Cambridge University Press: Cambridge, New York, 2001.10.1017/CBO9780511612626Search in Google Scholar

8. Schubert, K. The two-correlations model, a valence model for metallic phases. Struct. Bond 1977, 33, 139–177.10.1007/BFb0117580Search in Google Scholar

9. Lenz, J., Schubert, K. Über einige Leerstellen- und Stapelvarianten der Beta-Messing Strukturfamilie (On some vacancy and stacking variants of the beta-brass structure family; in German). Z. Metallkd. 1971, 62, 810–816. https://doi.org/10.1515/ijmr-1971-621110.Search in Google Scholar

10. Schubert, K. Crystal structures of beta brass like alloy phases. Trans. Jpn. Inst. Met. 1973, 14, 281–284. https://doi.org/10.2320/matertrans1960.14.281.Search in Google Scholar

11. Bradley, A. J., Thewlis, J. The structure of γ-brass. Proc. Roy. Soc. A 1926, 112, 678–692. https://doi.org/10.1098/rspa.1926.0134.Search in Google Scholar

12. Bradley, A. J., Taylor, A., Bragg, W. L. An X-ray analysis of the nickel-aluminium system. Proc. Roy. Soc. A 1937, 159, 56–72. https://doi.org/10.1098/rspa.1937.0056.Search in Google Scholar

13. Bradley, A. J., Taylor, A. XCIX. The crystal structures of Ni2Al3 and NiAl3. Phil. Mag. 1937, 23, 1049–1067. https://doi.org/10.1080/14786443708561875.Search in Google Scholar

14. Dong, C. The Al5Ni3 structure as approximant of quasicrystals. J. Phys. I France 1995, 5, 1625–1634. https://doi.org/10.1051/jp1:1995109.10.1051/jp1:1995109Search in Google Scholar

15. Kuntze, V., Lux, R., Hillebrecht, H. V2Cu3Ga8, Mo2Cu3Ga8 and W2Cu3Ga8—new compounds with a novel order variant of a bcc packing and motifs of self-similarity. J. Solid State Chem. 2007, 180, 198–206; https://doi.org/10.1016/j.jssc.2006.10.009.Search in Google Scholar

16. Lux, R., Kuntze, V., Hillebrecht, H. Synthesis and crystal structure of cubic V11Cu9Ga46 – a 512-fold super structure of a simple bcc packing. Solid State Sci. 2012, 14, 1445–1453. https://doi.org/10.1016/j.solidstatesciences.2012.07.028.Search in Google Scholar

17. Koffi, A., Ade, M., Hillebrecht, H. Synthesis, single crystal growth and crystal structure of Ta7Cu10Ga34 – a 8 × 4 × 2 super structure of CsCl. Z. Anorg. Allg. Chem. 2016, 642, 350–354. https://doi.org/10.1002/zaac.201500800.Search in Google Scholar

18. Giacovazzo, C., Monaco, H. L., Artioli, G., Viterbo, D., Milanesio, M., Gilli, G., Gilli, P., Zanotti, G., Ferraris, G., Catti, M. Fundamentals of Crystallography; Oxford University Press: Oxford, 2011.10.1093/acprof:oso/9780199573653.001.0001Search in Google Scholar

19. Hahn, T., Ed. International Tables for Crystallography, 5th ed., 1st online ed.; Springer Netherland: Berlin, 2007.10.1107/97809553602060000100Search in Google Scholar

20. Bhattacharya, K. Microstructure of Martensite: Why it Forms and How it Gives Rise to the Shape-Memory Effect? Oxford University Press: Oxford, UK, 2012.Search in Google Scholar

21. Koumatos, K., Muehlemann, A. Optimality of general lattice transformations with applications to the Bain strain in steel. Proc. Roy. Soc. A 2016, 472, 20150865. https://doi.org/10.1098/rspa.2015.0865.Search in Google Scholar PubMed PubMed Central

22. Chen, X., Song, Y., Tamura, N., James, R. D. Determination of the stretch tensor for structural transformations. J. Mech. Phys. Solids 2016, 93, 34–43. https://doi.org/10.1016/j.jmps.2016.02.009.Search in Google Scholar

23. Aizu, K. Determination of the state parameters and formulation of spontaneous strain for ferroelastics. J. Phys. Soc. Jpn. 1970, 28, 706–716. https://doi.org/10.1143/JPSJ.28.706.Search in Google Scholar

24. Momma, K., Izumi, F. Vesta 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. https://doi.org/10.1107/S0021889811038970.Search in Google Scholar

25. Palmer, D. C. Visualization and analysis of crystal structures using CrystalMaker software. Z. Kristallogr. 2015, 230, 559–572. https://doi.org/10.1515/zkri-2015-1869.Search in Google Scholar

26. Laurila, T., Vuorinen, V., Kivilahti, J. K. Interfacial reactions between lead-free solders and common base materials. Mater. Sci. Eng. R 2005, 49, 1–60. https://doi.org/10.1016/j.mser.2005.03.001.Search in Google Scholar

27. Schimpf, C., Kalanke, P., Shang, S. L., Liu, Z. K., Leineweber, A. Stacking disorder in metastable NiSn4. Mater. Des. 2016, 109, 324–333. https://doi.org/10.1016/j.matdes.2016.07.002.Search in Google Scholar

28. Leineweber, A., Wolf, C., Kalanke, P., Schimpf, C., Becker, H., Shang, S. L., Liu, Z. K. From random stacking faults to polytypes: a 12-layer NiSn4 polytype. J. Alloys Compd. 2019, 774, 265–273. https://doi.org/10.1016/j.jallcom.2018.09.341.Search in Google Scholar

29. Schmetterer, C., Flandorfer, H., Richter, K. W., Saeed, U., Kauffman, M., Roussel, P., Ipser, H. A new investigation of the system Ni–Sn. Intermetallics 2007, 15, 869–884. https://doi.org/10.1016/j.intermet.2006.10.045.Search in Google Scholar

30. Nowotny, H., Schubert, K. Die Kristallstruktur von Ni3Sn4. Naturwissenschaften 1944, 32, 76. https://doi.org/10.1007/BF01468013.Search in Google Scholar

31. Nowotny, H., Schubert, K. Die Kristallstruktur von Ni3Sn4. Z. Metallkd. 1946, 37, 23–31. https://doi.org/10.1515/ijmr-1946-371-204.Search in Google Scholar

32. Jeitschko, W., Jaberg, B. Structure refinement of Ni3Sn4. Acta Crystallogr. B 1982, 38, 598–600. https://doi.org/10.1107/S056774088200346X.Search in Google Scholar

33. Furuseth, S., Fjellvag, H., Marøy, K. Structural properties of Ni3+xSn4. Acta Chem. Scand. A 1986, 40, 695–700; https://doi.org/10.3891/acta.chem.scand.40a-0695.Search in Google Scholar

34. Bhan, S., Schubert, K. Zum Aufbau der Systeme Kobalt-Germanium, Rhodium-Silizium sowie einiger verwandter Legierungen. Z. Metallkd. 1960, 51, 327–339. https://doi.org/10.1515/ijmr-1960-510604.Search in Google Scholar

35. Richardson, M., Kierkegaard, P., Santesson, J., Holmberg, P., Eriksson, G., Blinc, R., Paušak, S., Ehrenberg, L., Dumanović, J. Crystal structure refinements of the B 20 and monoclinic (CoGe-type) polymorphs of FeGe. Acta Chem. Scand. 1967, 21, 753–760; https://doi.org/10.3891/acta.chem.scand.21-0753.Search in Google Scholar

36. Bhargava, M. K., Schubert, K. Kristallstruktur von NiSn. J. Less-Common Met. 1973, 33, 181–189. https://doi.org/10.1016/0022-5088(73)90037-4.Search in Google Scholar

37. Lihl, F., Kirnbauer, H. Untersuchung binärer metallischer Systeme mit Hilfe des Amalgamverfahrens. Das System Nickel-Zinn. Monatsh. Chem. 1955, 86, 745–751. https://doi.org/10.1007/BF00902566.Search in Google Scholar

38. Ma, Z. L., Xian, J. W., Belyakov, S. A., Gourlay, C. M. Nucleation and twinning in tin droplet solidification on single crystal intermetallic compounds. Acta Mater. 2018, 150, 281–294. https://doi.org/10.1016/j.actamat.2018.02.047.Search in Google Scholar

39. Yu, L. J., Yen, H. W., Wu, J. Y., Yu, J. J., Kao, C. R. Micromechanical behavior of single crystalline Ni3Sn4 in micro joints for chip-stacking applications. Mater. Sci. Eng., A 2017, 685, 123–130. https://doi.org/10.1016/j.msea.2017.01.004.Search in Google Scholar

40. Christian, J. W., Mahajan, S. Deformation twinning. Prog. Mater. Sci. 1995, 39, 1–157. https://doi.org/10.1016/0079-6425(94)00007-7.Search in Google Scholar

41. Weinberger, C. R., Boyce, B. L., Battaile, C. C. Slip planes in bcc transition metals. Int. Mater. Rev. 2013, 58, 296–314. https://doi.org/10.1179/1743280412Y.0000000015.Search in Google Scholar

42. Corby, R. N., Black, P. J. The structure of FeAl2 by anomalous dispersion methods. Acta Crystallogr. B 1973, 29, 2669–2677. https://doi.org/10.1107/S056774087300734X.Search in Google Scholar

43. Bastin, G. F., van Loo, F., Vrolijk, J., Wolff, L. R. Crystallography of aligned Fe-Al eutectoid. J. Cryst. Growth 1978, 43, 745–751. https://doi.org/10.1016/0022-0248(78)90155-0.Search in Google Scholar

44. Hirata, A., Mori, Y., Ishimaru, M., Koyama, Y. Role of the triclinic Al2Fe structure in the formation of the Al5Fe2-approximant. Phil. Mag. Lett. 2008, 88, 491–500. https://doi.org/10.1080/09500830802247151.Search in Google Scholar

45. Chumak, I., Richter, K. W., Ehrenberg, H. Redetermination of iron dialuminide, FeAl2. . Acta Crystallogr. C 2010, 66, i87–i88. https://doi.org/10.1107/S0108270110033202.Search in Google Scholar PubMed

46. Mihalkovic, M., Widom, M. Structure and stability of Al2Fe and Al5Fe2: first-principles total energy and phonon calculations. Phys. Rev. B 2012, 85, 14113. https://doi.org/10.1103/PhysRevB.85.014113.Search in Google Scholar

47. Scherf, A., Kauffmann, A., Kauffmann-Weiss, S., Scherer, T., Li, X., Stein, F., Heilmaier, M. Orientation relationship of eutectoid FeAl and FeAl2. J. Appl. Crystallogr. 2016, 49, 442–449. https://doi.org/10.1107/S1600576716000911.Search in Google Scholar PubMed PubMed Central

48. Li, L.-L., Su, Y., Beyerlein, I. J., Han, W.-Z. Achieving room-temperature brittle-to-ductile transition in ultrafine layered Fe–Al alloys. Sci. Adv. 2020, 6, eabb6658. https://doi.org/10.1126/sciadv.abb6658.Search in Google Scholar PubMed PubMed Central

49. Schmitt, A., Kumar, K. S., Kauffmann, A., Heilmaier, M. Microstructural evolution during creep of lamellar eutectoid and off-eutectoid FeAl/FeAl2 alloys. Intermetallics 2019, 107, 116–125; https://doi.org/10.1016/j.intermet.2019.01.015.Search in Google Scholar

50. Li, L., Beyerlein, I. J., Han, W. Interface-facilitated stable plasticity in ultra-fine layered FeAl/FeAl2 micro-pillar at high temperature. J. Mater. Sci. Technol. 2021, 73, 61–65; https://doi.org/10.1016/j.jmst.2020.09.018.Search in Google Scholar

51. Balanetskyy, S., Grushko, B., Velikanova, T. Y. Monoclinic Al2Fe phase, its equilibrium and nonequilibrium formation. Metallofiz. Noveishie Tekhnol. 2004, 26, 407–417.Search in Google Scholar

52. Ilatovskaia, M., Becker, H., Fabrichnaya, O., Leineweber, A. The η-Al5Fe2 phase in the Al–Fe system: the issue with the sublattice model. J. Alloys Compd. 2023, 936, 168361. https://doi.org/10.1016/j.jallcom.2022.168361.Search in Google Scholar

53. Gibson, J. S. K.-L., Pei, R., Heller, M., Medghalchi, S., Luo, W., Korte-Kerzel, S. Finding and characterising active slip systems: a short review and tutorial with automation tools. Materials 2021, 14, 407. https://doi.org/10.3390/ma14020407.Search in Google Scholar PubMed PubMed Central

54. Kamimura, Y., Edagawa, K., Takeuchi, S. Experimental evaluation of the Peierls stresses in a variety of crystals and their relation to the crystal structure. Acta Mater. 2013, 61, 294–309. https://doi.org/10.1016/j.actamat.2012.09.059.Search in Google Scholar

Received: 2023-05-25
Accepted: 2023-07-18
Published Online: 2023-08-01
Published in Print: 2023-09-26

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 28.4.2024 from https://www.degruyter.com/document/doi/10.1515/zkri-2023-0021/html
Scroll to top button