Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Differences and similarities in the chemical bonding of intermetallic phases in the Ca–Al–Pt system

Peter C. Müllera, Linda S. Reitza, Stefan Engelb, Richard Dronskowski*a and Oliver Janka*b
aInstitute of Inorganic Chemistry, RWTH Aachen University, Jülich-Aachen Research Alliance (JARA-CSD), RWTH Aachen University, 52056 Aachen, Germany
bInorganic Solid State Chemistry, Saarland University, Campus C4.1, 66123 Saarbrücken, Germany. E-mail: oliver.janka@uni-saarland.de

Received 24th April 2025 , Accepted 27th August 2025

First published on 28th August 2025


Abstract

Intermetallic compounds belong to an important class of materials, not only due to the sheer number of compounds known but also due to their application in everyday life. These compounds possess their very own peculiarities, especially when it comes to chemical bonding. To address this point, bonding analyses based on Crystal Orbital Bond Index (COBI) values, Löwdin charges, and – for the first time – ab initio oxidation numbers (ONai) were conducted, all extracted from delocalized plane-wave functions. From the integrated COBI values, to be understood as quantum-chemical bond orders, the differences and similarities in the bonding behavior of the elements, binary and ternary compounds in the Ca–Al–Pt system were analyzed. It became apparent that the Al–Pt interactions, almost regardless of the respective compounds, show significant covalency, while Ca–Al and Ca–Pt interactions are of ionic nature in most cases. Homoatomic Al–Al or Pt–Pt interactions, however, tend to be ambivalent, depending on the respective crystal structure of a given compound.


1. Introduction

Alloys and intermetallic compounds are important materials used in a manifold of modern everyday applications. Light-weight alloys, usually based on Be, Mg, Al, and Ti, are utilized in consumer goods, as well as in the transportation and construction sector.1–7 In addition, the most commonly used permanent magnets,8,9 materials with high thermal stability and corrosion resistance,10–13 or heterogeneous catalysts can be found amongst intermetallic compounds.14–19 They achieved practical use already way before their intrinsic characteristics were fully understood since they can often be obtained directly from the reaction of the constituent metals by metallurgical processes, arc-melting or metal fluxes.20–22 Nonetheless, the fundamental understanding of crystal chemistry, existence ranges and the chemical bonding in this class of materials is still ongoing.23–27 To quote Yuri Grin: “The main problem of chemists with intermetallic compounds is that they do not follow the usual valence rules. Therefore, for a longer time, these substances were not really considered as inorganic compounds”.28

Disobeying the classical valence rules leads to some problems because unless the underlying wave function is known and analyzed the nature of the chemical bond cannot be understood.29–38 What are the contributions to a certain interaction between two atoms? Ionic interactions or rather covalency? Does a concept of, say, electronegativity still apply? When going to the extremes, e.g., Cs2Pt39,40 or CsAu,40–42 transparent compounds with salt-like behavior can be observed; however, when ΔEN is not as striking as in these examples, is there still a polarity in these intermetallic phases? The term ‘polar intermetallics’ is used quite frequently to describe compounds where a certain polarization within the structure is either assumed or proven. The question of how to properly address the bonding in solids dates back to almost a century: Laves43 and Pauling44–46 introduced atomic packing and “resonance” concepts for crystalline solids, while Hume-Rothery47 addressed the electronic (metallic) state and critical valence-electron concentrations.

Science has come a long way, since nowadays chemical bonding48 may be addressed computationally based on density-functional theory, either resting on the wave function49–51 or on the electron density.52–56 The main difference between both approaches lies in the examined quantity:34 a density-based method such as Bader's quantum theory of atoms in molecules (QTAIM) semi-classically partitions the electron density according to its topology. Even though this looks attractive since the density (1) is easily accessible from DFT calculations and (2) also experimentally observable from, e.g., X-ray diffraction, the density lacks essential information, which is crucial for a well-grounded evaluation. As the density relates to the absolute square of the wave function, the phase information (i.e., the sign of the wave function) is completely lost in the density, thus we cannot distinguish between bonding and antibonding, at least not from the wave function.

In a solid-state context, there is yet another hindrance: common DFT programs employ a delocalized plane-wave basis lacking local (chemical) information that would require atom-centered basis functions. In order to extract these bonding data, the program LOBSTER57 conducts a so-called projection from a plane-wave onto an atomic-orbital basis. Therefore, we regain chemistry in terms of, e.g., Löwdin charges58 and crystal orbital bond index (COBI) values, the latter including the phase information.59 Both established tools allow for a bond classification in terms of ionicity and covalency. Additionally, in this recent contribution, we also introduce wave-function derived oxidation numbers that aim to resolve the schism between quantum-mechanical charges and empirical oxidation states.

The case study in this paper will be the elements Ca, Al and Pt, selected binaries of the systems Ca–Al, Ca–Pt and Al–Pt, as well as all reported ternary compounds of the entire Ca–Al–Pt system. Fig. 1 depicts a Gibbs triangle showing all reported binary and ternary phases based on the Pearson database60 and the recent reports. Besides the crystal structures of the three elements Ca, Al and Pt,61–63 CaAl2 (ref. 64) and CaAl4,64 and CaPt2 (ref. 65) as well as Al2Pt,66 and AlPt67 were selected from the respective binary phase diagrams. Furthermore, CaAlPt, Ca2AlPt2 and CaAl2Pt were selected as representatives of ternary compounds. Our essential goal is to quantify the bonding nature in these intermetallic compounds, namely (1) to check if interatomic distances can be used as an identifier for bonding or non-bonding scenarios and (2) to identify similarities amongst the crystallographically significantly different compounds and highlight specific peculiarities that arise based on these quantum-chemical calculations.


image file: d5sc02993g-f1.tif
Fig. 1 Gibbs triangle showing all reported phases in the ternary system Ca–Al–Pt (yellow). Element symbols are shown in black; binary compounds in the Ca–Al system are depicted in grey, in the Ca–Pt system in blue and in the Al–Pt system in red.

2. Computational methods

Electronic structure calculations for the elements and selected binary as well as ternary compounds were performed using the projector augmented wave method (PAW) of Blöchl,68 as implemented in the Vienna ab initio simulation package (VASP).69–73 VASP calculations employed the Ca_sv, Al, Pt_pv, H and N pseudopotentials. The calculations started from the experimental crystallographic data but allowing for full structural optimizations, including lattice parameters and atomic positions. These were conducted for all compounds including electronic structure analyses. In all calculations, correlation and exchange were treated by the generalized gradient approximation of Perdew, Burke, and Ernzerhof (GGA-PBE).74 The cutoff energy for the plane-wave calculations was chosen to be a high value of 800 eV, and the convergence criteria for the energy differences between two iterative steps were set to 10−6 and 10−5 eV for the electronic and ionic steps, respectively. Brillouin zone integration was carried out using a k-point mesh with a spacing of ≈0.02 Å−1 for all compounds.

All corresponding electronic structures, based on the optimized calculations, were projected from plane waves onto a local orbital basis using the LOBSTER (Local Orbital Basis Suite Towards Electronic-structure Reconstruction) program package.50,57,75–77 By this method, the local density-of-states matrices (i.e., energy-resolved wave-function eigenvectors) become available, which enable the calculation of the crystal orbital bond index (COBI),59 a generalized solid-state molecular bond index introduced by Wiberg78 and Mayer,79 as well as gross populations and atomic charges in a Löwdin-style formalism, in addition to ab initio oxidation numbers.58

3. Results and discussion

Before we proceed to analyze the chemical bonding, let us reiterate the underlying methods to describe a chemical bond. On a qualitatively correct level, (organic) chemists draw lines – “paired” electrons in a two-center two-electron bond – between two carbon atoms in order to symbolize an attractive C–C bond keeping these atoms together. Quantum-mechanically, this interaction can be described by a plethora of quantities. One of the most useful descriptors may be the bond index originally envisioned by Wiberg and Mayer that directly translates into the Lewis bond order, i.e., 1 for a single bond, 2 for a double bond and so on. This bond index was then generalized to the solid state by means of the crystal orbital bond index (COBI) linking the Wiberg–Mayer bond idea to solid-state descriptors such as the crystal orbital overlap population by Hughbanks and Hoffmann49 and the crystal orbital Hamilton population by Dronskowski and Blöchl.50,75 The energy integral of the COBI, dubbed ICOBI, equals the bond order between two atoms in a solid. In molecular chemistry, integer bond orders prevail, but there are molecular exceptions such as a bond order of 1/2 in the hydrogen-molecule cation, H2+, or in the benzene aromatic C–C bond with a bond order of 3/2. In (inter-)metallic phases, however, fractional bond orders are the rule, typically characterized by ICOBI numbers well below the value of a single bond. This can be directly attributed to the larger coordination numbers found in condensed matter.

3.1. Oxidation numbers from quantum chemistry

Having introduced covalent bonding in solids, it is time to introduce a novel, yet related, method to calculate ab initio oxidation numbers (ONai) such as to expand LOBSTER's quantum-chemical toolkit. All currently available population analyses homolytically split the bonding electrons, on purpose, and attribute half to each contributing atom: in a Mulliken-style scheme, this property is easily seen in the symmetric partitioning of the overlap population between two orbitals μ and ν. The Löwdin population analysis, however, uses an orthogonal basis, so an overlap population does not exist in the first place, but the homolytic splitting of bonding electrons is also utilized. This property becomes apparent when we consider the idempotency of the density matrix P, i.e., Tr(P) = ½Tr(|P|2), and it lets us partition the electron density of any given compound into atom-centered electrons (½Pμμ2) as well as bond-centered electrons (½|Pμν|2). Just like Mulliken, Löwdin assigns the same number of bonding electrons to each atom without consideration of element-specific properties, such as electronegativity.

In the present work, we do address this issue, however, by introducing a weighting factor that allocates the bonding electrons to a more electronegative atom, in the spirit of heterolytic electron partitioning. This extension of the Löwdin population analysis then yields ab initio oxidation numbers ONai that are defined as the difference between the number of electrons in the neutral atom Ne,A and the population after heterolytic bond splitting:

 
image file: d5sc02993g-t1.tif(1)

As the electronegativity is not straightforwardly accessible from quantum-mechanical calculations, we derive it by referring to the Hamilton matrix elements Hμν and constructing the weighting factor wμν:

 
wμν = 1 + erf(α(HννHμμ)) (2)

The error function and the parameter α determine the “smearing” of the electronic partitioning between two atoms, and it ensures that minute energy differences between Hμμ and Hνν of, say, 0.1 eV as well as numerical noise do not cause an unreasonably drastic electron transfer. In the present study, we chose a value of α = 10 eV−1 which, for an energy difference of 0.1 eV, leads to the assignment of 92% electron density to the more electronegative atom and the remaining 8% to the less electronegative atom. A transfer of more than 99% is then performed at a difference of 0.165 eV.

The algorithm formulated in eqn (1) and (2) is visualized in Fig. 2 for the simple example of molecular ammonia, NH3. Fig. 2a shows the calculation of formal charges that result from a homolytic splitting of bonds. As such, the nitrogen atom is left with five electrons, and each hydrogen atom keeps one electron, leading to all atoms being formally neutral. When calculating oxidation numbers (cf. Fig. 2b), all N–H bonds are split heterolytically, so the bonding electrons are assigned to the more electronegative bonding partner, nitrogen in this case. This ionic limit then leads to oxidation numbers of N−III and H+I.


image file: d5sc02993g-f2.tif
Fig. 2 Algorithm to calculate (a and c) formal charges and (b and d) oxidation numbers from (a and b) empirical Lewis formulae and (c and d) ab initio population analyses for ammonia. Orbital and bond populations are given as numbers. A homolytic bond splitting, as done for the calculation of formal charges and α = 0, leads to the assignment of half the bonding electrons to each bonding partner. In the case of oxidation numbers, a heterolytic bond splitting is performed by α ≫ 0 such that all bonding electrons are assigned to the more electronegative atom.

Going from classical Lewis formulae to quantum chemistry, we can quantify the number of electrons located on atoms and bonds by the population analyses already introduced above. As can easily be seen in Fig. 2c, setting the weight factor α to zero, splitting the bonding electrons (indicated by solid red and blue lines) is symmetric, so in case of, say, the 1s–2px/y bond, both bonding partners receive half of the 0.59 bonding electrons, 0.29 each. In sum, this separation leads to Löwdin populations and charges of −0.91 for nitrogen and +0.30 for hydrogen.

If the weighting factor α is chosen to be larger, the partitioning of the bonding population becomes heterolytical. As the 1s orbital of hydrogen is significantly higher in energy compared to the valence orbitals of nitrogen, the electrons are shifted strictly towards nitrogen, in line with the empirical expectation from electronegativities. For the 1s–2px/y bond, the total 0.59 bonding electrons are thus shifted completely to nitrogen, mirroring the recipe of empirical oxidation numbers. In the end, the ONai derived from this algorithm (H+0.75 and N−2.25) match the empirical numbers a lot closer than the respective Löwdin charges. At this point, it should be noted that fractional ONai are an inherent feature of the population analyses used in the calculus. While integer oxidation numbers are a consequence of the classical derivation, this does not apply to quantum chemistry. Analogously, Mulliken/Löwdin charges and populations as well as bond orders by means of ICOBI have fractional values in the vast majority of cases, especially in intermetallic phases as presented in the following.

3.2. Similarities and differences

Before we go into detail, we want to give an overview of all compounds included in our study, as summarized in Fig. 3. Starting with covalent bonding, Fig. 3a sets the ICOBI of each contact in relation to the respective interatomic distance. In this depiction, several trends are apparent: (1) bond orders steadily decrease with increasing interatomic distance, even though outliers are apparent. While individual elements and crystal structures affect the relationship, the overall trend is clearly visible, especially inside certain bond types. (2) Al–Al bonds are longer than Al–Pt bonds with the same bond strength, while Pt–Pt bonds are significantly weaker. Caused by the relative sizes of Al and Pt valence orbitals, this was already found for individual compounds, but the present summary shows this observation on a much wider scale. (3) Interactions involving Ca do not contain significant covalency. In the given compounds, the Ca atoms act as electron donors and can be formally identified as Ca2+, unable to form any covalent bonds due to its closed-shell (4s0) configuration.
image file: d5sc02993g-f3.tif
Fig. 3 (a) Bond length plotted versus the ICOBI per bond. (b) Ab initio oxidation number plotted versus Löwdin charge.

The ionic bonding analyses in this study are conducted based on Löwdin charges and, additionally, using the newly introduced ab initio oxidation numbers ONai. Their relation is shown in Fig. 3b. In all examined cases, Ca purely serves as an electron donor with charges and oxidation numbers close to the ideal +2. For Al and Pt, simple expectations are not met, however. Based on electronegativities of Al and Pt, one would assume (partially) cationic Al and anionic Pt. Fig. 3b reveals that all of the examples in our examination possess negatively charged aluminum and positively charged platinum atoms, in contrast to electronegativities of Al and Pt, respectively, 1.61 and 2.28 on Pauling's scale.80 Since Löwdin charges do not reflect the energetic orbital arrangement and symmetrically attribute the density of bonding electrons to both bonding partners, charges do not trivially turn into oxidation states. This disagreement is solved by introducing a weighting scheme based on atomic orbital energies (see the explanation above), ultimately leading to the formulation of ab initio oxidation numbers. Although these oxidation numbers are often fractional, not integer like the classical counterpart, the overall fit is significantly improved. The majority of Al atoms reside in the top left part of Fig. 3b and Pt atoms in the bottom right part. This picture can be traced back to an increased bonding electron transfer from Al to Pt that is revealed when going from Löwdin charges to oxidation numbers. As most of the systems provided in Fig. 3 follow the expected course of bond strength vs. bond length, we will focus the following more detailed discussion of our results on the most interesting members of the Ca–Al–Pt family. For reference, data on all compounds are summarized in the SI.

3.3. The elements

All three elements crystallize in their stable allotropes at room temperature and ambient pressure, i.e., in the cubic crystal system with the face centered space group Fm[3 with combining macron]m, adopting the Cu type structure (Fig. 4a).61–63 The DFT-optimized interatomic distances are 284 pm for Pt, 286 pm for Al, and 393 pm for Ca. In all three cases, these distances are longer compared to the sum of the covalent radii (Pt: 258 pm; Al: 250 pm; Ca: 348 pm (ref. 80)), not too surprising taking into account the larger coordination numbers.
image file: d5sc02993g-f4.tif
Fig. 4 (a) Unit cell of elemental Ca, Al and Pt (Cu type, Fm[3 with combining macron]m) depicted with the corresponding coordination environment. Unit cells of (b) cubic Ca, (c) cubic Pt and (d) cubic Al are presented with Löwdin charges, ab initio oxidation numbers (ONai), interatomic distances and respective ICOBI values.

Based on the interatomic distances, one would expect rather weak covalent interactions to the neighbors, which can easily be verified using the ICOBI values of the nearest-neighbor bonds: Al–Al has the largest value of 0.23, Ca–Ca (0.12) and Pt–Pt (0.09) are smaller, indicating even weaker covalent bonding in these metals.81 If we sum up all of the respective bond indices, we arrive at the total bond capacity of an atom that equals its valence, similar to the empirical bond valence sum.82 Including only the 12 nearest-neighbor interactions, the valences are 2.76 for Al, 1.44 for Ca, and 1.08 for Pt.

This relation of rather small metal–metal bond orders and valences can easily be rationalized: if we consider elemental calcium, each Ca atom directly bonds with 12 nearest neighboring atoms, meaning that two valence electrons per Ca atom are distributed over all of these bonding partners. Consequently, the expected number of “shared” electrons between two Ca atoms will be around 2/12 ≈ 0.17, corresponding to a Ca–Ca bond order of ca. 0.08, as mirrored by ICOBI.

Naturally, none of these compounds show any electron transfer as all Löwdin charges are ±0. This agrees with the ab initio oxidation numbers (ONai) that all have a value of ±0, a trivial consequence of an element.

3.4. Binary phases

As for the binary phases, stable and, where possible, multiple-times reported (according to the Pearson database60) representatives from each binary system were selected: CaAl2 (ref. 83) and CaAl4 (ref. 84) for the Ca–Al system, CaPt2 (ref. 65) for the Ca–Pt system and finally Al2Pt and AlPt for the Al–Pt system.

CaAl2 and CaPt2 both adopt the cubic Laves phase (MgCu2 type; Fd[3 with combining macron]m).85 Here, Al4/Pt4 tetrahedra are found, which are connected via all four corners to a network. The Ca atoms reside in cavities of the said framework (Fig. 5a and b). The Al–Al distances in CaAl2 are 281 pm and therefore are well in line with elemental Al.


image file: d5sc02993g-f5.tif
Fig. 5 Unit cells of (a) cubic CaAl2 and (b) CaPt2, both of MgCu2 type (Fd[3 with combining macron]m), and (c) monoclinic CaAl4 (CaGa4 type, C2/m). Ca, Al and Pt atoms are shown in light grey, dark grey and black circles, respectively. The Löwdin charges, ab initio oxidation numbers (ONai), interatomic distances and respective ICOBI values for the different structures are also provided.

Despite the similar interatomic distance, the ICOBI (0.43) of this contact is larger by a factor of two, and this is directly related to the coordination number of six, as compared to CN = 12 in the fcc structure. Speaking of ionicity, one would expect the formation of aluminides, which is directly reflected in significant Löwdin charges. The Ca atoms exhibit positive charges with +1.54 for CaAl2, and Al anions are formed with a charge of −0.77. This charge transfer is even more pronounced when looking at the ONai that are +1.86 and −0.93, corresponding to Ca+II and Al−I. A recent review on binary alkaline-earth trielides utilizing the Bader formalism observed a similar charge transfer.86

In isostructural CaPt2, the Pt–Pt distances are 270 pm and therefore are significantly shorter compared to elemental Pt, suggesting covalent bonding interaction. Indeed, the same picture is evident as found in CaAl2: covalently bonded Pt–Pt and rather ionic Ca–Pt interactions. Yet, there are certain differences addressed in the following: while the formation of cationic Ca and anionic Pt is given by both Löwdin charges and ONai, the quantitative charge transfer does not match the empirical expectations. Based on electronegativity, one would assume more negative Pt compared to Al, but both the Löwdin charge and ONai are smaller than the respective values of Al in CaAl2.

From a quantum-mechanical point of view, however, elemental Al (3s2 3p1) possesses a less than half-filled valence shell, while the valence of Pt (6s2 5d8) is more than half-filled. Following the (generalized) octet rule and correlation arguments, adding an electron to Pt is energetically less favorable than adding an electron to Al. This destabilization is also visible in the (I)COBI of the Pt–Pt contacts. Comparing elemental Pt and CaPt2, the ICOBI is larger for the shorter contact in the binary compound, but only by a small amount – especially in comparison with Al/CaAl2. Pushing electrons from Ca onto Pt leads to the population of antibonding levels in the Pt–Pt interactions, thereby weakening the individual bonds.

CaAl4 adopts a monoclinic structure (CaGa4 type, C2/m) that can be derived from the tetragonal BaAl4 type structure by a group-subgroup formalism.64 Once again, the Al atoms form a network with the Ca atoms residing in cavities (Fig. 5c). Here, one would naively consider the shorter Al–Al contacts to be covalent bonding interactions, while the longer ones would be considered non-bonding. Interestingly, the rather long Al1–Al1 contacts still exhibit significant ICOBI values of about 0.23. The Al1–Al2 distances are significantly shorter, leading to an increase in the ICOBI values to slightly above the half bond order. The shortest distance, finally, is the Al2–Al2 interaction between the layered fragments with a stunningly large ICOBI value of 0.81, indicating a very strong covalent bond, approaching the classical single bond order. The Löwdin charges finally reflect the coordination environments: while Al1 exhibits in a wider sense a coordination number of 12 (Al1@Al24Al14Ca4), Al2 has a coordination number of 9 (Al2@Al2Al14Ca4) with overall shorter distances and therefore a higher overall Löwdin charge. This trend has also been observed by charge transfer analyses based on the definition by Bader.64,86 Interestingly, the ONai further differentiate between both Al sites resulting in an anionic Al2 (−1.27) and a cationic Al1 (+0.31), solid-state disproportionation, so to speak. In order to understand this behavior, a fundamental solid-state periodic property needs to be recalled, namely the Madelung field. Pure electrostatics leads to an attractive (=stabilizing) force between cations and anions. Considering the rather short Al2–Ca distance relative to Al1–Ca, it is safe to assume a stronger stabilization of Al2, in turn lowering the orbital energies of Al2, anion-like. Thus, the bonding electrons of the Al1–Al2 bonds move to the more electronegative Al2 atoms, and this ultimately leads to the formation of Al1 cations and Al2 anions, at least formally.

Finally, the compounds from the Al–Pt system should be addressed (Fig. 6). When looking at the bond indices in Al2Pt (Fig. 6a), the Al–Pt contacts are significantly shorter compared to the Al–Al interactions, even though both exhibit similar ICOBI values of 0.27 and 0.26. The ionic nature of this compound, as given by the Löwdin charges, does not reflect the empirical expectation from electronegativities. This tentative disagreement can immediately be solved by the respective oxidation numbers ONai. Using orbital energies as criteria and heterolytic charge allocation, bonding electrons of the Al–Pt bonds are transferred from Al 3p to the lower-lying Pt 5d orbitals. This way, the positively charged Pt (+0.26) adopts an anionic oxidation number (−0.82) and the negatively charged Al (−0.13) has a cationic oxidation number (+0.41). Such a charge transfer of 0.54 electrons per Al atom to Pt is only revealed by the combination of an atom-centered (ionic) and a bond-centered (covalent) analysis, as explained above.


image file: d5sc02993g-f6.tif
Fig. 6 Unit cells of (a) cubic Al2Pt (CaF2 type, Fm[3 with combining macron]m), and (b) cubic AlPt (FeSi type, P213). Al and Pt atoms are shown in dark grey and black circles, respectively. The Löwdin charges, ab initio oxidation numbers (ONai), interatomic distances and respective ICOBI values for the different structures are also provided.

In AlPt (Fig. 6b), a contrasting picture is observed. There are three distinct Al–Pt contacts, suggesting similar bonding based on the distances. The first two interactions indeed have identical ICOBI values of 0.25, and the third interaction, which is almost identical in length, only has an ICOBI value of 0.17. This puzzling difference shall be investigated using the energy-dependent COBI, as shown in Fig. S2. At a quick glance, we identify antibonding levels directly below the Fermi level, and their amount is largest in the weakest Al–Pt bond, so the small ICOBI is not due to decreased orbital interaction but a shift from occupied bonding to antibonding levels. Note that this destabilizing part is counteracted by the remaining Al–Pt bonds that have a multiplicity of three in contrast to the “weakest” bond appearing only once per polyhedron. On a descriptive level, we may infer that the remaining contacts are stabilized at the cost of the weak bond resulting in a net stabilizing effect. Ionicity, on the other hand, is more transparent. As also found for Al2Pt above, the Löwdin charges show counter-intuitive Al anions and Pt cations. This discrepancy in terms of electronegativities is again resolved by the oxidation numbers being significantly smaller than the Löwdin charges. This time, however, the oxidation numbers are almost neutral with ±0.05, and they do not mirror distinct ions.

In general, it can be observed that the charge transfer from Al to Pt increases with increasing Al content. This is in line with recent studies on the Al–Pt system based on QTAIM,87,88 although the absolute values reported in literature are larger than our Löwdin charges and ONai.

3.5. Ternary phases

In the following paragraphs, the crystal structures of some of the compounds listed in Table S5 will be discussed exemplarily alongside similarities, differences and specific features that will be highlighted by the bonding analysis later.

We start with equiatomic CaAlPt89,90 which crystallizes in the orthorhombic crystal system with the space group Pnma and adopts the TiNiSi type structure. The crystal structure can be described based on a network formed by the Al and Pt atoms with the Ca atoms residing in cavities (Fig. 7a). The Al–Pt distances are relatively short within the network; additionally, elongated Al–Al distances can be observed; however, no Pt–Pt interactions are present. When compared with the distance discussions above, at least Al–Pt bonds are present. Since the Al to Pt ratio is 1[thin space (1/6-em)]:[thin space (1/6-em)]1, forming Al–Pt interactions is a necessity, while Al–Al and Pt–Pt interactions are not necessarily required. The Ca–Al and Ca–Pt distances are relatively short, but clearly distinct to at least the Al–Pt contacts and in the range of the sum of covalent radii (vide infra). Based on the different distances, one can interpret this structure as a polyanionic [AlPt]δ network with Caδ+ cations residing in the cavities. This is a frequently observed picture, especially in the Al-rich compounds.


image file: d5sc02993g-f7.tif
Fig. 7 Unit cells of (a) orthorhombic CaAlPt (TiNiSi type, Pnma) and (b) monoclinic Ca2AlPt2 (Ca2SiIr2 type, C2/c) with corresponding Löwdin charges, ab initio oxidation numbers (ONai), and ICOBI values for selected bonds. Ca, Al and Pt atoms are shown in light grey, white and black circles, respectively.

The Löwdin charges underline the picture of a polyanionic network (Ca: +1.26; Al: −1.08; Pt: −0.17); however, as already observed for the binary compounds in the Al–Pt system, i.e., the Al atoms carry a higher negative charge counterintuitive to the electronegativities. In this case, both the Al and the Pt atoms are negatively charged and also show an anionic oxidation number. The ONai, however, are in better agreement with the trend derived from the electronegativities and arrive at −1.01 for Pt while Al yields −0.65. The interatomic Al–Pt distances are in a similar range as the ones observed in the binaries, but overall higher ICOBI values are observed. This can be attributed to the additional electron transfer from the Ca atoms onto the polyanion, filling bonding levels. Interestingly, the longest Al–Pt contact exhibits the strongest covalency (0.43). Besides the Al–Pt contacts also one relatively long Al–Al distance is present. This interaction would not be considered bonding with respect to the distances found in elemental Al or the sum of the covalent radii, but the ICOBI value of 0.26 clearly indicates bonding interactions.

We now draw our attention to monoclinic Ca2AlPt2 (Fig. 7b, Ca2SiIr2 type). Here, there are chains of Pt atoms with alternating shorter and longer distances. These chains run parallel to the ab plane with an angle of 63.6° between them. The Al atoms connect two of the chains with rather short Al–Pt contacts, always bridging the longer Pt–Pt distances. The Ca atoms reside in the created cavities. This compound was reported by Doverbratt and coworkers who also analyzed the bonding.91 The authors highlight the linear platinum chains that distort pairwise into shorter dumbbells and longer Pt⋯Pt contacts. When looking at the ICOBI values (Fig. 7b), 0.30 can be found for the short Pt–Pt distance, while only 0.03 is calculated for the long distance. This is in line with Doverbratt et al.;91 however, their bond orders were 0.48 and 0.26 according to the non-quantum-chemical bond-length bond-strength approach by Brese and O'Keeffe.82 They state that “the interactions within the anionic substructures are essentially nonbonding (or slightly bonding) and indicate that the electrostatic repulsions are suppressed when replacing the anionic Ge bridging elements by cationic Al atoms”. This can be clearly seen by the ICOBI values of 0.36 for the Al–Pt interactions. The ionic bonds in Ca2AlPt2 suggest a similar decomposition as found in the previously discussed CaAlPt. Ca is cationic, and Al and Pt form an anionic network with Al being slightly more negatively charged than Pt. As found before, the oxidation numbers reveal a polarization of the Al–Pt bonds that results in more negative Pt and less negative Al.

When finally going to CaAl2Pt (orthorhombic, MgAl2Cu type, Cmcm, Fig. 8a),92 a compound with Al being the majority element is discussed, so can distinct Al–Al bonding be observed? When only focusing on the arrangement of the Al atoms, we see corrugated honeycomb layers all in boat conformation with the Pt atoms residing in the center of each hexagon. As described before, the Al–Pt contacts are the shortest ones observed in the structure, while the Al–Al distances are slightly longer. This generates [Al2Pt] layers which are separated by the Ca atoms leading to Ca–Pt distances of 314 pm and Ca–Al distances of 323 and 342 pm. The shortest distance between the layers not involving Ca atoms is Al–Al = 349 pm. From a crystal-chemical point of view, one would discard this distance as not being involved in any bonding. Interestingly, the same distance elongates to 415 pm in SrAl2Pt92 and even further to 471 pm in BaAl2Pt.92 While CaAl2Pt is quite stable towards moisture, SrAl2Pt and BaAl2Pt decompose rapidly, so does the short Al–Al distance stabilize CaAl2Pt? In the rare-earth representatives, this contact decreases down to 281 pm for ScAl2Pt,93 which is well in line with what is considered a bonding interaction based on structural considerations. In the relaxed DFT model (Fig. 8a), the distance in CaAl2Pt is slightly shorter (340 pm); however, the ICOBI value is 0.31, that is about a third of a single bond. This value is especially remarkable since the shorter Al–Al contact with a length of 292 pm has a similar ICOBI of 0.34 despite the significant bond-length difference, which addresses a considerable covalency to the longer interaction. Talking of ionicity, the Löwdin charges suggest an anionic Al–Pt network that contains Ca cations. The positive sign of the Pt charge (+0.22) turns into an anionic ONai (−0.18) that is primarily caused by a shift of bonding electrons from the Ca–Pt bonds to Pt. Al has the same values for the Löwdin charge and oxidation number.


image file: d5sc02993g-f8.tif
Fig. 8 Unit cell of (a) orthorhombic CaAl2Pt (MgCuAl2 type, Cmcm) and (b) hexagonal Ca2Al3Pt (MgZn2/Mg2Cu3Si type, P63/mmc). The Löwdin charges, ab initio oxidation numbers (ONai), interatomic distances and corresponding ICOBI values for different bonds are indicated. Ca, Al and Pt atoms are shown in light grey, dark grey and black circles, respectively.

Another interesting Al–Al interaction can be observed in Ca2Al3Pt (Fig. 8b). Here, the Al atoms form a 63 Kagome net with the Pt atoms connecting two layers. Within the Al layer, there are two different interactions, the shorter one being 272 and the longer being 285 pm. Puzzling, the shorter interaction exhibits the smaller ICOBI value of 0.37, while the significantly longer distance shows a significantly higher covalency, with 0.52 being the highest ICOBI value for an Al–Al interaction observed. The lower ICOBI for the shorter bond can be traced back to (occupied) antibonding contribution, below the Fermi energy. Note that a shorter bond length may increase orbital overlap, but this also holds for antibonding interactions that decrease the bond strength as in the present example. Why, however, does this unexpected course of bond strength vs. bond length appear in the first place for a compound with only one crystallographic site per element? To answer, we take a quick look at the crystal structure: the triangles formed by the longer Al–Al bonds are capped by a calcium atom (dashed lines in Fig. 8b), essentially forming a trigonal pyramid; this structural feature does not exist for the shorter Al–Al contacts. It is safe to assume that the cationic presence of Ca2+ stabilizes the bonds in the anionic Al-network, thus leading to an increased ICOBI despite the longer Al–Al distance. Charges and ONai of Ca2Al3Pt match our findings from the previous discussion.

4. Conclusions

In this work, we presented a thorough investigation of the bonding properties for members of the Ca–Al–Pt family starting with the elements, then continuing via binary to ternary compounds. As already found in previous work, the chemical bonding in intermetallic phases is not necessarily analogous to simple molecular compounds due to complexities of the solid state and metallicity. Using the toolbox provided by LOBSTER, we investigated covalent bonding by the crystal orbital bond index and ionic bonding by means of Löwdin charges and ab initio oxidation numbers. By comparison of both descriptors for ionicity, we demonstrated the strength of the oxidation numbers that directly relate to oxidation states covered in any basic chemistry course. The ability to calculate such numbers promises unexpected insights into intermetallic phases where classical concepts do not apply.

Author contributions

All authors have accepted responsibility for the entire content of this submitted manuscript and approved the submission.

Conflicts of interest

The authors declare no conflicts of interest regarding this article.

Data availability

Crystallographic data can be found in the primary articles cited throughout the manuscript, further information is given in the SI. If any data is missing, it can be requested from the corresponding authors upon reasonable request. See DOI: https://doi.org/10.1039/d5sc02993g.

Acknowledgements

Funding was provided by the Deutsche Forschungsgemeinschaft DFG (grant JA 1891-10-1).

References

  1. D. Altenpohl, Aluminium und Aluminiumlegierungen, Springer, Berlin, Heidelberg, Germany, 1965 Search PubMed.
  2. C. Kammer, Aluminium-Taschenbuch 1: Grundlagen und Werkstoffe, Aluminium-Zentrale: Düsseldorf, Germany, 1995 Search PubMed.
  3. F. Ostermann, Anwendungstechnologie Aluminium. Springer Vieweg: Berlin, Heidelberg, Germany, 2014 Search PubMed.
  4. O. Janka, Metallic light-weight alloys: Al, Ti, Mg, in Applied Inorganic Chemistry, ed. Pöttgen, R., Jüstel, T. and Strassert, C., De Gruyter, Berlin, Germany, 2022, pp. 158–173 Search PubMed.
  5. M. Buchner and O. Janka, Volume 1 From Construction Materials to Technical Gases – Chapter 2.7 Be and Be alloys, in Applied Inorganic Chemistry, ed. Pöttgen, R., Jüstel, T. and Strassert, C. A., De Gruyter, Berlin, 2023, pp. 221–228 Search PubMed.
  6. S. Engel and O. Janka, Volume 1 From Construction Materials to Technical Gases - Chapter 2.10 Shape memory alloys, in Applied Inorganic Chemistry, ed. Pöttgen, R., Jüstel, T. and Strassert, C. A., De Gruyter, 2023, pp. 255–264 Search PubMed.
  7. S. Engel and O. Janka, Volume 1 From Construction Materials to Technical Gases - Chapter 2.11 Bulk metallic glasses, in Applied Inorganic Chemistry, ed. Pöttgen, R., Jüstel, T. and Strassert, C. A., De Gruyter, 2023, pp. 265–270 Search PubMed.
  8. K. Strnat, G. Hoffer, J. Olson, W. Ostertag and J. J. Becker, A Family of New Cobalt-Base Permanent Magnet Materials, J. Appl. Phys., 1967, 38, 1001–1002 CrossRef.
  9. J. J. Croat, J. F. Herbst, R. W. Lee and F. E. Pinkerton, High-energy product Nd-Fe-B permanent magnets, Appl. Phys. Lett., 1984, 44, 148–149 CrossRef.
  10. A. Rahmel, M. Schütze and W. J. Quadakkers, Fundamentals of TiAl oxidation – A critical review, Mater. Corros., 1995, 46, 271–285 CrossRef.
  11. G. Kresse, M. Schmid, E. Napetschnig, M. Shishkin, L. Köhler and P. Varga, Structure of the Ultrathin Aluminum Oxide Film on NiAl(110), Science, 2005, 308, 1440–1442 CrossRef PubMed.
  12. P. K. Datta, H. L. Du, J. S. Burnell-Gray and R. E. Ricker, Corrosion of Intermetallics, in Corrosion: Materials, ASM International, Materials Park, OH, USA, 2005, vol. 13B Search PubMed.
  13. P. Jozwik, W. Polkowski and Z. Bojar, Applications of Ni3Al Based Intermetallic Alloys – Current Stage and Potential Perceptivities, Materials, 2015, 8, 2537–2568 CrossRef.
  14. M. Armbrüster, K. Kovnir, M. Behrens, D. Teschner, Y. Grin and R. Schlögl, Pd–Ga Intermetallic Compounds as Highly Selective Semihydrogenation Catalysts, J. Am. Chem. Soc., 2010, 132, 14745–14747 CrossRef PubMed.
  15. M. Armbrüster, K. Kovnir, M. Friedrich, D. Teschner, G. Wowsnick, M. Hahne, P. Gille, L. Szentmiklósi, M. Feuerbacher, M. Heggen, F. Girgsdies, D. Rosenthal, R. Schlögl and Y. Grin, Al13Fe4 as a low-cost alternative for palladium in heterogeneous hydrogenation, Nat. Mater., 2012, 11, 690–693 CrossRef PubMed.
  16. I. Antonyshyn, O. Sichevych, K. Rasim, A. Ormeci, U. Burkhardt, S. Titlbach, S. A. Schunk, M. Armbrüster and Y. Grin, Chemical behaviour of CaAg2 under ethylene epoxidation conditions, Eur. J. Inorg. Chem., 2018, 2018, 3933–3941 CrossRef.
  17. K. L. Hodge and J. E. Goldberger, Transition Metal-Free Alkyne Hydrogenation Catalysis with BaGa2, a Hydrogen Absorbing Layered Zintl Phase, J. Am. Chem. Soc., 2019, 141, 19969–19972 CrossRef PubMed.
  18. J. Li and S. Sun, Intermetallic Nanoparticles: Synthetic Control and Their Enhanced Electrocatalysis, Acc. Chem. Res., 2019, 52, 2015–2025 CrossRef PubMed.
  19. M. Armbrüster, Intermetallic compounds in catalysis – a versatile class of materials meets interesting challenges, Sci. Technol. Adv. Mater., 2020, 21, 303–322 CrossRef PubMed.
  20. M. G. Kanatzidis, R. Pöttgen and W. Jeitschko, The metal flux: A preparative tool for the exploration of intermetallic compounds, Angew. Chem., Int. Ed., 2005, 44, 6996–7023 CrossRef PubMed.
  21. W. Steurer and J. Dshemuchadse, Intermetallics. Union of Crystallography, Oxford University Press, Oxford, 2016 Search PubMed.
  22. R. Pöttgen and D. Johrendt, Intermetallics - Synthesis, Structure, Function, De Gruyter, Berlin, Boston, 2019 Search PubMed.
  23. A. R. Oganov, Modern Methods of Crystal Structure Prediction, Wiley-VCH, Weinheim, 2010 Search PubMed.
  24. J. George and G. Hautier, Chemist versus Machine: Traditional Knowledge versus Machine Learning Techniques, Trends Chem., 2021, 3, 86–95 CrossRef.
  25. D. Dahliah, G. Brunin, J. George, V.-A. Ha, G.-M. Rignanese and G. Hautier, High-throughput computational search for high carrier lifetime, defect-tolerant solar absorbers, Energy Environ. Sci., 2021, 14, 5057–5073 RSC.
  26. J. George, G. Petretto, A. Naik, M. Esters, A. J. Jackson, R. Nelson, R. Dronskowski, G.-M. Rignanese and G. Hautier, Automated Bonding Analysis with Crystal Orbital Hamilton Populations, ChemPlusChem, 2022, 87, e202200123 CrossRef PubMed.
  27. A. M. Ganose, H. Sahasrabuddhe, M. Asta, K. Beck, T. Biswas, A. Bonkowski, J. Bustamante, X. Chen, Y. Chiang, D. C. Chrzan, J. Clary, O. A. Cohen, C. Ertural, M. C. Gallant, J. George, S. Gerits, R. E. A. Goodall, R. D. Guha, G. Hautier, M. Horton, T. J. Inizan, A. D. Kaplan, R. S. Kingsbury, M. C. Kuner, B. Li, X. Linn, M. J. McDermott, R. S. Mohanakrishnan, A. N. Naik, J. B. Neaton, S. M. Parmar, K. A. Persson, G. Petretto, T. A. R. Purcell, F. Ricci, B. Rich, J. Riebesell, G.-M. Rignanese, A. S. Rosen, M. Scheffler, J. Schmidt, J.-X. Shen, A. Sobolev, R. Sundararaman, C. Tezak, V. Trinquet, J. B. Varley, D. Vigil-Fowler, D. Wang, D. Waroquiers, M. Wen, H. Yang, H. Zheng, J. Zheng, Z. Zhu and A. Jain, Atomate2: modular workflows for materials science, Digital Discovery, 2025, 4, 1944–1973 RSC.
  28. Y. Grin, Crystal Structure and Bonding in Intermetallic Compounds, in Comprehensive Inorganic Chemistry II, ed. Reedijk, J. and Poeppelmeier, K., Elsevier, Amsterdam, 2nd edn, 2013, pp. 359–373 Search PubMed.
  29. J. K. Burdett, Perspectives in structural chemistry, Chem. Rev., 1988, 88, 3–30 CrossRef.
  30. R. Nesper, Chemische Bindungen - intermetallische Verbindungen, Angew. Chem., 1991, 103, 805–834 CrossRef.
  31. R. Nesper, Bonding Patterns in Intermetallic Compounds, Angew. Chem., Int. Ed., 1991, 30, 789–817 CrossRef.
  32. J. K. Burdett, Chemical Bonds: A Dialog, John Wiley & Sons Ltd, Chichester, England, 1997 Search PubMed.
  33. T. A. Albright, J. K. Burdett and M.-H. Whangbo, Orbital Interactions in Chemistry, Wiley & Sons Ltd, Hoboken, New Jersey, 2013 Search PubMed.
  34. G. J. Miller, Y. Zhang and F. R. Wagner, Chemical Bonding in Solids, in Handbook of Solid State Chemistry, Wiley-VCH, Weinheim, 2017, pp. 405–489 Search PubMed.
  35. V. Smetana, M. Rhodehouse, G. Meyer and A.-V. Mudring, Gold Polar Intermetallics: Structural Versatility through Exclusive Bonding Motifs, Acc. Chem. Res., 2017, 50, 2633–2641 CrossRef CAS PubMed.
  36. Q. Lin and G. J. Miller, Electron-poor polar intermetallics: Complex structures, novel clusters, and intriguing bonding with pronounced electron delocalization, Acc. Chem. Res., 2018, 51, 49–58 CrossRef.
  37. R. O. Jones, The chemical bond in solids – revisited, J. Phys.: Condens. Matter, 2022, 34, 343001 CrossRef.
  38. S. Steinberg, Tellurides as Zintl phases?, J. Phys.: Condens. Matter, 2025, 37, 273001 CrossRef PubMed.
  39. A. Karpov, J. Nuss, U. Wedig and M. Jansen, Cs2Pt: A Platinide(-II) Exhibiting Complete Charge Separation, Angew. Chem., Int. Ed., 2003, 42, 4818–4821 CrossRef PubMed.
  40. M. Jansen, Effects of relativistic motion of electrons on the chemistry of gold and platinum, Solid State Sci., 2005, 7, 1464–1474 CrossRef.
  41. W. J. Peer and J. J. Lagowski, Metal-ammonia solutions. 11. Au, a solvated transition metal anion, J. Am. Chem. Soc., 1978, 100, 6260–6261 CrossRef.
  42. M. Jansen, The chemistry of gold as an anion, Chem. Soc. Rev., 2008, 37, 1826–1835 RSC.
  43. F. Laves, XVI. Die Bau-Zusammenhänge innerhalb der Kristallstrukturen, Z. Kristallogr., 1930, 73, 202–265 CAS.
  44. L. Pauling, The principles determining the structure of complex ionic crystals, J. Am. Chem. Soc., 1929, 51, 1010–1026 CrossRef.
  45. L. Pauling, Atomic Radii and interatomic distances in metals, J. Am. Chem. Soc., 1947, 69, 542–553 CrossRef.
  46. L. Pauling, The nature of the chemical bond and the structure of molecules and crystals: an introduction to modern structural chemistry, Cornell University Press, Ithaca, NY, USA, 1960 Search PubMed.
  47. W. Hume-Rothery, The metallic state: electrical properties and theories. Clarendon Press, Oxford, Great Britai, 1931 Search PubMed.
  48. R. Dronskowski, Chemical Bonding, Walter De Gruyter, Berlin/Boston, 2023 Search PubMed.
  49. T. Hughbanks and R. Hoffmann, Chains of trans-edge-sharing molybdenum octahedra: metal-metal bonding in extended systems, J. Am. Chem. Soc., 1983, 105, 3528–3537 CrossRef.
  50. R. Dronskowski and P. E. Blöchl, Crystal orbital Hamilton populations (COHP): energy-resolved visualization of chemical bonding in solids based on density-functional calculations, J. Phys. Chem., 1993, 97, 8617–8624 CrossRef.
  51. R. Nelson, C. Ertural, P. C. Müller and R. Dronskowski, Chemical bonding with plane waves, in Comprehensive Inorganic Chemistry III, ed. Reedijk, J. and Poeppelmeier, K. R., Elsevier, Oxford, 3rd edn, 2023, pp. 141–201 Search PubMed.
  52. R. F. W. Bader, A quantum theory of molecular structure and its applications, Chem. Rev., 1991, 91, 893–928 CrossRef.
  53. A. Savin, R. Nesper, S. Wengert and T. F. Fässler, ELF: The Electron Localization Function, Angew. Chem., Int. Ed., 1997, 36, 1808–1832 CrossRef.
  54. D. C. Fredrickson, DFT-chemical pressure analysis: Visualizing the role of atomic size in shaping the structures of inorganic materials, J. Am. Chem. Soc., 2012, 134, 5991–5999 CrossRef PubMed.
  55. G. Frenking and S. Shaik, The chemical bond: fundamental aspects of chemical bonding, John Wiley & Sons, Weinheim, Germany, 2014 Search PubMed.
  56. F. R. Wagner and Y. Grin, Chemical bonding analysis in position space, in Comprehensive Inorganic Chemistry III, ed. Reedijk, J. and Poeppelmeier, K. R., Elsevier, Oxford, 3rd edn, 2023, pp. 222–237 Search PubMed.
  57. R. Nelson, C. Ertural, J. George, V. L. Deringer, G. Hautier and R. Dronskowski, LOBSTER: Local orbital projections, atomic charges, and chemical-bonding analysis from projector-augmented-wave-based density-functional theory, J. Comput. Chem., 2020, 41, 1931–1940 CrossRef.
  58. C. Ertural, S. Steinberg and R. Dronskowski, Development of a robust tool to extract Mulliken and Löwdin charges from plane waves and its application to solid-state materials, RSC Adv., 2019, 9, 29821–29830 RSC.
  59. P. C. Müller, C. Ertural, J. Hempelmann and R. Dronskowski, Crystal Orbital Bond Index: Covalent Bond Orders in Solids, J. Phys. Chem. C, 2021, 125, 7959–7970 CrossRef.
  60. P. Villars and K. Cenzual, Pearson's Crystal Data: Crystal Structure Database for Inorganic Compounds, release 2024/25, ASM International®, Materials Park, Ohio, USA, 2024 Search PubMed.
  61. A. W. Hull, A new method of X-ray crystal analysis, Phys. Rev., 1917, 10, 661–696 CrossRef.
  62. A. W. Hull, The positions of atoms in metals, Trans. Am. Inst. Electr. Eng., 1919, 38, 1445–1466 Search PubMed.
  63. A. W. Hull, The Arrangement of Atoms in Some Common Metals, Science, 1920, 52, 227–229 CrossRef.
  64. S. Engel, E. C. J. Gießelmann, L. E. Schank, G. Heymann, K. Brix, R. Kautenburger, H. P. Beck and O. Janka, Theoretical and 27Al NMR spectroscopic investigations of binary intermetallic alkaline-earth aluminides, Inorg. Chem., 2023, 62, 4260–4271 CrossRef.
  65. E. A. Wood and V. B. Compton, Laves-phase compounds of alkaline earths and noble metals, Acta Crystallogr., 1958, 11, 429–433 CrossRef.
  66. E. Zintl, A. Harder and W. Haucke, Legierungsphasen mit Fluoritstruktur (22. Mitteilung über Metalle und Legierungen), Z. Phys. Chem., 1937, 35, 354–362 CrossRef.
  67. K. Schubert, W. Burkhardt, P. Esslinger, E. Günzel, H. G. Meissner, W. Schütt, J. Wegst and M. Wilkens, Einige strukturelle Ergebnisse an metallischen Phasen, Naturwissenschaften, 1956, 43, 248–249 CrossRef.
  68. P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef.
  69. G. Kresse and J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558–561 CrossRef.
  70. G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef PubMed.
  71. G. Kresse and J. Furthmüller, Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
  72. G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758–1775 CrossRef CAS.
  73. G. Kresse, M. Marsman and J. Furthmüller, Vienna Ab-initio Simulation Package VASP: the Guide, Computational Materials Physics, Faculty of Physics, Universität Wien, Vienna, Austria, 2014 Search PubMed.
  74. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation made simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef.
  75. V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, Crystal Orbital Hamilton Population (COHP) Analysis As Projected from Plane-Wave Basis Sets, J. Phys. Chem. A, 2011, 115, 5461–5466 CrossRef CAS PubMed.
  76. S. Maintz, V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, Analytic projection from plane-wave and PAW wavefunctions and application to chemical-bonding analysis in solids, J. Comput. Chem., 2013, 34, 2557–2567 CrossRef.
  77. S. Maintz, V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, LOBSTER: A tool to extract chemical bonding from plane-wave based DFT, J. Comput. Chem., 2016, 37, 1030–1035 CrossRef.
  78. K. B. Wiberg, Application of the pople-santry-segal CNDO method to the cyclopropylcarbinyl and cyclobutyl cation and to bicyclobutane, Tetrahedron, 1968, 24, 1083–1096 CrossRef.
  79. I. Mayer, Charge, bond order and valence in the AB initio SCF theory, Chem. Phys. Lett., 1983, 97, 270–274 CrossRef.
  80. J. Emsley, The Elements, Clarendon Press, Oxford University Press, Oxford, New York, 1998 Search PubMed.
  81. L. S. Reitz, J. Hempelmann, P. C. Müller, R. Dronskowski and S. Steinberg, Bonding Analyses in the Broad Realm of Intermetallics: Understanding the Role of Chemical Bonding in the Design of Novel Materials, Chem. Mater., 2024, 36, 6791–6804 CrossRef.
  82. N. E. Brese and M. O’Keeffe, Bond-Valence Parameters for Solids, Acta Crystallogr., 1991, B47, 192–197 Search PubMed.
  83. H. Nowotny, E. Wormnes and A. Mohrnheim, Untersuchungen in den Systemen Aluminium-Kalzium, Magnesium-Kalzium und Magnesium-Zirkon, Z. Metallkd., 1940, 32, 39–42 Search PubMed.
  84. G. J. Miller, F. Li and H. F. Franzen, The structural phase transition in calcium-aluminum compound (CaAl4): a concerted application of Landau theory and energy band theory, J. Am. Chem. Soc., 1993, 115, 3739–3745 CrossRef.
  85. E. C. J. Gießelmann, R. Pöttgen and O. Janka, Laves phases: superstructures induced by coloring and distortions, Z. Anorg. Allg. Chem., 2023, 649, e202300109 CrossRef.
  86. H. P. Beck, A review on binary alkaline-earth trielides: an cverview of compositions and structures, of their stabilities and energetics and of criteria for the validity of the Zintl-concept, Z. Anorg. Allg. Chem., 2024, 650, e202400011 CrossRef.
  87. A. Baranov, M. Kohout, F. R. Wagner, Y. Grin and W. Bronger, Spatial chemistry of the aluminium–platinum compounds: a quantum chemical approach, Z. Kristallogr., 2007, 222, 527–531 CrossRef.
  88. I. Antonyshyn, O. Sichevych, U. Burkhardt, A. M. Barrios Jiménez, A. Melendez-Sans, Y.-F. Liao, K.-D. Tsuei, D. Kasinathan, D. Takegami and A. Ormeci, Al–Pt intermetallic compounds: HAXPES study, Phys. Chem. Chem. Phys., 2023, 25, 31137–31145 RSC.
  89. P. Kenfack Tsobnang, D. Fotio, S. Ponou and C. Fon Abi, Calcium platinum aluminium, CaPtAl, Acta Crystallogr., 2011, E67, i55 Search PubMed.
  90. F. Hulliger, On new ternary aluminides LnPdAl and LnPtAl, J. Alloys Compd., 1993, 196, 225–228 CrossRef.
  91. I. Doverbratt, S. Ponou, Y. Zhang, S. Lidin and G. J. Miller, Linear Metal Chains in Ca2M2X (M = Pd, Pt; X = Al, Ge): Origin of the Pairwise Distortion and Its Role in the Structure Stability, Chem. Mater., 2015, 27, 304–315 CrossRef.
  92. F. Stegemann, T. Block, S. Klenner, Y. Zhang, B. P. T. Fokwa, A. Timmer, H. Mönig, C. Doerenkamp, H. Eckert and O. Janka, From 3D to 2D: structural, spectroscopical and theoretical investigations of the dimensionality reduction in the [PtAl2]δ polyanions of the isotypic MPtAl2 series (M = Ca–Ba, Eu), Chem.–Eur. J., 2019, 25, 10735–10747 CrossRef PubMed.
  93. M. Radzieowski, F. Stegemann, C. Doerenkamp, S. F. Matar, H. Eckert, C. Dosche, G. Wittstock and O. Janka, Correlations of crystal and electronic structure via NMR and XPS spectroscopy in the RETMAl2 (RE = Sc, Y, La–Nd, Sm, Gd–Tm, Lu; TM = Ni, Pd, Pt) series, Inorg. Chem., 2019, 58, 7010–7025 CrossRef PubMed.

Footnotes

We note that the entire DFT procedure leading to the orbital picture is from first principles, without empirical parameters. To arrive at the chemists’ somewhat arbitrary heterolytical bond splitting, a likewise arbitrary parameter α cannot be avoided such as to not cause unreasonable shifts of bonding electrons.
Note that we chose covalent radii here on purpose. While the metallic radii exactly match the interatomic distances by definition, we want to contextualize covalent bonding properties that are better described by covalent radii as their sum corresponds to the length of a single bond that has ICOBI = 1.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.