Open Access Article
Fathima Thanha Thazhe Namboorikandy
and
Pattiyil Parameswaran
*
Department of Chemistry, National Institute of Technology Calicut, Kozhikode, Kerala 673601, India. E-mail: param@nitc.ac.in
First published on 20th November 2025
A comprehensive geometrical and bonding analysis of heterobimetallic dimetallocenes of the form [Cp–Be–M–Cp]+, where M represents group 13 elements (B, Al, Ga, In, Tl), is presented. The equilibrium geometry of dimetallocenes indicates a collinear geometry along the centre of the two Cp rings, the Be centre and the group 13 element, except for the Tl complex. Optimised geometries reveal that Be–C bond distances decrease when the group 13 element changes from B to Tl. The molecular orbital and NBO analyses show a Be–M σ-bond formed by the overlap of the 2s orbital of Be with the spn hybrid orbital of the group 13 element. The percentage of s character in the spn hybrid orbital increases as the group 13 element changes from B to Tl. However, the natural population analysis (NPA) indicates a significant charge polarisation with a low value of the Wiberg bond index. Energy decomposition analysis with natural orbitals for chemical valence (EDA-NOCV) reveals that the Be–M σ bond can be represented by a donor–acceptor interaction (Cp–Be+ ← M–Cp). Apart from the σ bonding interaction, there exists a slight π bonding interaction arising from hyperconjugative donation from the [BeCp]+ bonding molecular orbital to the [BCp] antibonding molecular orbital. Even though the strength of the Be–M bond decreases, the percentage of covalent interaction increases when the group 13 element changes from B to Tl.
The unexpected discovery of decamethyldizincocene (Cp*–Zn–Zn–Cp*, Cp* = C5Me5)26 by Carmona and coworkers in 2004 opened a new dimension in metallocene chemistry. In this D5h symmetric compound, a pair of Zn atoms is sandwiched between two Cp* rings, and the Zn–Zn bond is aligned collinearly with the C5 axes of the Cp* rings. Carmona also synthesised a similar dizincozene, Zn2(η5-C5Me4Et)2,27 with a substituted cyclopentadienyl ring. Kress28–32 confirmed from the electronic structural studies that the Zn atom has a +1 oxidation state rather than the usual +2 oxidation state for group 12 metals. Later, theoretical studies by many scientists concluded that the Zn–Zn bond is a single σ bond supported by weak π-type d–d and p–p interactions.33,34 Later, other bimetallic metallocenes of group-12 elements, including Zn,35,36 Cd37,38 and Hg,39 were also theoretically predicted (Scheme 1(ii)).
The discovery of transition metal dimetallocenes paved the way for main-group dimetallocenes. Schaefer and co-workers40 theoretically predicted the existence of homobimetallic dimetallocenes of alkaline-earth metals based on the view that both Zn and alkaline-earth metals possess an ns2 valence shell electronic configuration. Similarly, the electronic structure of stable binuclear metallocenes of group 2 metals has also been studied.41 Frenking and coworkers subsequently studied multimetallocenes with Be, Mg, Ca, and Zn.42 These studies showed that Cp2Be2 (or Cp*2Be2) is more stable than Cp*2Zn2. This led to the groundbreaking discovery of unusual metal–metal bonding among the main-group homobimetallic dimetallocenes. Boronski, Aldridge, and co-workers experimentally characterised a diberyllocene with a Be–Be single bond.43 They also isolated complexes with the first Be–Zn, Be–Al,44 Be–In and Be–Ga45 bonds. Later, many other complexes with homobimetallic46 and heterobimetallic47 bonding involving Be and alkaline earth metals have been reported. The metal–metal bonding in homonuclear dimetallocenes of Be, Mg, and Ca has been investigated in a theoretical study of A2N-MM′-NA2 (A = H, Li, Na)48 complexes. However, dimetallocenes of p-block elements are still unknown. Despite numerous attempts to isolate a silicon dimetallocene, all of them failed, as the alleged decamethyldisilicocene disproportionates into silicon(0) and decamethylsilicocene (Cp*2Si).49,50 Main-group bis-element sandwiches formed by donor–acceptor interactions of MCp and ECp (M = Li, Na, K; E = B, Al, Ga) have been studied by Timoshkin and Schaefer.51 Previous theoretical work52 and recent experimental studies53 have shown that heterobimetallic dimetallocenes with donor–acceptor interactions can preferentially adopt double-decker structures, such as [CpBe⋯CpE]+, rather than the conventional dimetallocene form [CpBe ← ECp]+. Recently, a lithium–aluminium dimetallocene with a highly polar Li–Al bond was isolated by Schafer in 2024.54 It is significantly reinforced by dispersion interactions, which arise between the two highly substituted Cp ligands. Li and co-workers elucidated the bonding characteristics in (η5-C5H5)2M2 complexes through computational analysis.55 Meanwhile, Gosch and Wilson56 expanded this understanding by examining related heterobimetallic systems, thereby providing a comparative perspective on metal–metal and metal–ligand interactions across main-group and transition metals.
Unlike classical transition metal metallocenes, such as Cp2Fe, which exhibit significant covalent bonding due to d-orbital participation, metallocenes of alkaline earth metals, like Cp2Ba, are predominantly ionic, reflecting pronounced differences in bonding nature and reactivity. However, the bonding in heterobimetallic dimetallocenes incorporating both beryllium and group 13 elements remains largely unexplored. At the same time, similar systems with two of the same metals or with group 1 and group 13 metals have been studied. Notably, beryllium is expected to favour electrostatic interactions, whereas group 13 metals are known for electron sharing and multicenter bonding. This research gap leaves open the question of how these differing bonding tendencies interact within a single molecule. We hypothesise that such systems may exhibit unique donor–acceptor interactions, offering new insights into bonding beyond the established electrostatic and covalent paradigms in metallocene chemistry.
The Be–C bond distances decrease as the group 13 element changes from B to Tl. This can be attributed to the donation of electrons from the bonding molecular orbital corresponding to Cp and Be interactions in [CpBe]+ to the antibonding molecular orbital corresponding to Cp and M interactions in [MCp] (vide infra). All metal–metal interactions can be considered as single bonds, where these distances are slightly shorter than single bond covalent radii proposed by Pyykkö and Atsumi, with values of 1.02 Å (Be), 0.85 Å (B), 1.26 Å (Al), 1.24 Å (Ga), 1.42 Å (In), and 1.44 Å (Tl).57 Although certain values are derived from a limited sample size (notably for Be), they provide a valuable comparison of metal–metal bonding. The Be–M (M = Al, Ga, In) bond distances observed in our complexes were compared with experimentally reported values. The experimentally known Be–Al bond lengths are 2.474 Å and 2.432 Å,44 while the calculated Be–Al distance in [Cp–Be–Al–Cp]+ is 2.279 Å. The calculated value closely matches the expected value for a typical single covalent bond. The Be–Ga (2.206 Å)44 and Be–In (2.358 Å)45 bond lengths determined experimentally align well with our calculated values of 2.213 Å and 2.364 Å, respectively.
Except for the Be–Tl dimetallocene, all other dimetallocenes have almost similar metal–ring (M–C) bond distances, which suggests that the hapticity of the cyclopentadienyl ring is 5. Cp–Be–Tl–Cp has one shorter Tl–C bond of 2.373 Å, two Tl–C bonds of 2.555 Å and two longer bonds of 2.791 Å. Therefore, the geometrical parameters suggest the bonding possibilities of η1–η3.
Our primary focus is to study the nature of metal–metal bonding in heterobimetallic dimetallocenes of Be. Molecular orbital analysis indicates that all dimetallocenes have occupied the Be–M σ-bonding orbital, viz HOMO−2 for [Cp–Be–B–Cp]+ and [Cp–Be–Al–Cp]+ and HOMO−4 for other dimetallocenes (Fig. 2). The molecular orbitals corresponding to Be–M π bonds are vacant. However, the valence p orbitals of Be and group 13 elements are involved in bonding interaction with π-molecular orbitals of the cyclopentadienyl ring, as indicated by the frontier occupied molecular orbitals of the system. All the MOs of the five dimetallocenes of Be are given in the SI. All the molecules exhibit a significant HOMO–LUMO energy gap, indicating their greater stability. The aforementioned bonding description is corroborated by the numerical data obtained from the natural bond orbital (NBO) analysis (Table 1). The Be atom in all dimetallocenes shows a positive value, which lies in the range of 1.27–1.53e. The positive charge on Be is moreover consistent with the electronegativity of the attached group 13 element, where the Be atom shows the highest positive charge in [Cp–Be–B–Cp]+. The charge analysis suggests that Be is electrophilic. However, it should be noted that previously reported beryllium–aluminyl complexes by Boronski and Aldridge44 suggested a nucleophilic Be(δ−)–Al(δ+) bond. This variation is attributed to the difference in the electronegativity of the atom/group attached to Al. The atoms/groups coordinated to aluminium in the beryllium–aluminyl complex are two N atoms and one O atom, whereas the atom/group attached to Al in [Cp–Be–Al–Cp]+ is a C atom. The charge of the Cp ring is negative except for that of the Cp ring attached to B. This indicates a significant electrostatic interaction between Be and the group 13 element, as indicated by the relative contribution of the orbital towards the Be–M bond (Table 2). The Wiberg bond index (WBI) values are less than one, indicating charge polarisation in the Be–M bond. The charge difference between Be and group 13 elements is in the order of Be–B (1.78e) ≫ Be–Tl (0.77e) > Be–Ga (0.74e) > Be–In (0.71e) > Be–Al (0.50e). The calculated WBI values are correlated with the above-mentioned charge differences. The Be–M σ-bond is formed by the overlap of the 2s orbital of Be (90.37%–94.20%) with the spn hybrid orbital of the group 13 element, M. The percentage of s character in the spn hybrid orbital increases as the group 13 element changes from B (55.78%) to In (84.7%). Therefore, the orbital overlap increases as the group 13 element changes from B to Tl in [Cp–Be–M–Cp]+. The NBO calculation reveals a lone pair on Tl, with a major contribution from the 6s orbital (94.2%), having an occupancy of 1.72e (Table 2). The η1 to η3 hapticity of the Cp ring connected to Tl can be attributed to the larger atomic size as well as the predominant relativistic effect of the 6s orbital of Tl.58
![]() | ||
| Fig. 2 Important frontier molecular orbitals of dimetallocenes at the M06/def2-TZVPP//BP86/def2-TZVP level of theory; eigenvalues are given in eV in parentheses; isosurface value = 0.03. | ||
| [Cp–Be–M–Cp]+ | Wiberg bond index (Be–M) | qCp (attached to Be) | qBe | qM | qCp (attached to M) |
|---|---|---|---|---|---|
| [Cp–Be–B–Cp]+ | 0.475 | −0.41 | 1.531 | −0.245 | +0.14 |
| [Cp–Be–Al–Cp]+ | 0.710 | −0.71 | 1.271 | 0.764 | −0.33 |
| [Cp–Be–Ga–Cp]+ | 0.613 | −0.70 | 1.352 | 0.614 | −0.27 |
| [Cp–Be–In–Cp]+ | 0.592 | −0.71 | 1.354 | 0.639 | −0.29 |
| [Cp–Be–Tl–Cp]+ | 0.588 | −0.71 | 1.364 | 0.591 | −0.25 |
| Be–M | Occupancy | Contribution from Be | Contribution from M | Relative contribution (Be : M) |
||
|---|---|---|---|---|---|---|
| s-Orbital contribution | p-Orbital contribution | s-Orbital contribution | p-Orbital contribution | |||
| Be–B | 1.99 | 94.20% | 5.79% | 5.79% | 44.07% | 15.3 : 84.6 |
| Be–Al | 1.99 | 91.64% | 8.14% | 8.14% | 24.45% | 25.8 : 74.1 |
| Be–Ga | 2.00 | 90.37% | 9.54% | 9.54% | 16.25% | 20.8 : 79.1 |
| Be–In | 2.00 | 90.51% | 9.21% | 9.21% | 15.12% | 19.9 : 80.0 |
| Be–Tl | — | — | — | — | — | — |
To understand the quantitative nature of the above-mentioned bonding description, we have performed an energy decomposition analysis coupled with natural orbital for chemical valence (EDA-NOCV) for all heterobimetallic dimetallocenes [Cp–Be–M–Cp]+. The choice of electronic states for the fragments is crucial in the EDA-NOCV analysis. Two different intuitive bonding possibilities have been considered for each dimetallocene (1–5, Scheme 2 and Table 3). The bonding representation A shows a donor–acceptor interaction with [CpBe]+ and [CpM] fragments. The bonding possibility B is an electron-sharing interaction with [CpBe] and [CpM]+. The best bonding situation has the lowest numerical value for the orbital interaction term, ΔEorb. The lowest orbital interaction energy, ΔEorb, indicates how closely the electronic states of the fragments match with those in the molecule.59 Since the ΔEorb value for bonding representation A is relatively low compared to that of B for all dimetallocenes, the former representation can be considered as the best bonding representation. It is noteworthy that the ΔEPauli for the bonding representation B is much higher than that in bonding representation A. The preparation energy (ΔEprep) for bonding possibility A is lower than for the other possibilities studied. Hence, bonding possibility A is considered for further analysis. The ΔEint can be considered as the measure of the strength of the Be–M bond, which increases as the group 13 element changes from B to Tl. The ΔEint correlates well with the dissociation energy of the Be–M bond as well. The electrostatic interaction energy, ΔEelect, increases and Pauli's repulsion energy, ΔEPauli, decreases as M changes from B to Tl. On the other hand, the variation in orbital stabilisation energy, ΔEorb, is minimal. Among the attractive components of interaction energy, the percentage of orbital stabilisation energy, ΔEorb, increases (38.64 to 90.16%) and the percentage of electrostatic energy, ΔEelect, decreases (61.35% to 9.83%) when the group 13 element moves down the group (Fig. 3). The Be–M σ-bond is formed by the donation of a lone pair of electrons from the spn hybrid orbital of the group 13 element to the vacant 2s orbital of Be (Fig. 4). The charge difference between [CpBe]+ and [MCp] and the electronegativity difference between Be and the group 13 element, M, is of the order Be–B ≫ Be–Ga > Be–In > Be–Tl > Be–Al, which accounts for the increased covalent character as the group 13 element changes from B to Tl, except for [Cp–Be–Al–Cp]+.
![]() | ||
| Fig. 4 Orbital interaction diagram of the [Cp–Be–B–Cp]+ dimetallocene showing frontier valence MOs. The fragments are [CpBe]+ (left side) and [BCp]0 (right side). | ||
| Be–B | Be–Al | Be–Ga | Be–In | Be–Tl | Be–B | |||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| A | B | A | B | A | A | B | A | B | A | A | B | |
| a Values in parentheses give the percentage contribution to ΔEorb + ΔEelect.b Percentage contribution to ΔEorb.c ΔErest = ΔEorb − (ΔEσ + ΔEπ⊥ + ΔEπ∥).d Dissociation energy (De). | ||||||||||||
| ΔEint | −120.45 | −146.43 | −78.78 | −121.80 | −66.69 | −150.68 | −59.05 | −134.28 | −49.46 | −105.90 | −120.45 | −146.43 |
ΔEelect a |
−101.67 (61.35%) | −101.40 (24.20%) | −35.69 (35.48%) | −80.70 (25.60%) | −25.30 (29.89%) | −114.28 (30.66%) | −15.32 (20.74%) | −124.42 (35.80%) | −6.08 (9.83%) | −147.39 (43.29%) | −101.67 (61.35%) | −101.40 (24.20%) |
| ΔEPauli | 50.35 | 277.60 | 26.08 | 197.70 | 22.54 | 226.54 | 19.60 | 217.97 | 17.67 | 239.80 | 50.35 | 277.60 |
ΔEorb a |
−64.03 (38.64%) | −317.52 (75.79%) | −64.89 (64.51%) | −234.53 (74.39%) | −59.34 (70.10%) | −258.35 (69.33%) | −58.53 (79.25%) | −223.03 (64.19%) | −55.75 (90.16%) | −193.02 (56.70%) | −64.03 (38.64%) | −317.52 (75.79%) |
| ΔEdisp | −5.10 | −5.10 | −4.28 | −4.28 | −4.59 | −4.59 | −4.80 | −4.80 | −5.30 | −5.30 | −5.10 | −5.10 |
ΔEσ b |
−49.17 (76.79%) | −259.81 (81.82%) | −57.81 (89.08%) | −186.14 (79.36%) | −50.97 (85.89%) | −129.96 (50.30%) | −50.54 (86.34%) | −122.41 (54.88%) | −47.76 (85.66%) | −60.56 (31.37%) | −49.17 (76.79%) | −259.81 (81.82%) |
ΔEπ⊥ b |
−3.99 (6.23%) | −31.33 (9.86%) | −1.83 (2.82%) | −34.08 (14.53%) | −2.65 (4.46%) | −90.44 (35.01%) | −2.43 (4.15%) | −61.94 (27.77%) | −2.66 (4.77%) | −93.58 (48.48%) | −3.99 (6.23%) | −31.33 (9.86%) |
ΔEπ∥ b |
−3.99 (6.23%) | −10.35 (3.25%) | −1.83 (2.82%) | −2.65 (1.12%) | −2.65 (4.46%) | −20.98 (8.12%) | −2.43 (4.15%) | −26.08 (11.69%) | −2.11 (3.78%) | −11.85 (6.13%) | −3.99 (6.23%) | −10.35 (3.25%) |
ΔErest c |
−6.88 (10.74%) | −16.03 (5.04%) | −3.42 (5.27%) | −11.66 (4.97%) | −3.07 (5.17%) | −16.97 (6.56%) | −3.13 (5.34%) | −12.60 (5.64%) | −3.22 (5.77%) | −27.08 (14.02%) | −6.88 (10.74%) | −16.03 (5.04%) |
| ΔEprep | 11.93 | 37.91 | 11.29 | 54.31 | 26.96 | 110.95 | 18.91 | 94.14 | 11.05 | 67.49 | 11.93 | 37.91 |
−De d |
−108.52 | −108.52 | −67.49 | −67.49 | −39.73 | −39.73 | −40.14 | −40.14 | −38.41 | −38.41 | −108.52 | −108.52 |
The orbital stabilisation energy (ΔEorb) can be categorised into three symmetry-adapted components: one σ-interaction and two π-interactions (π∥ and π⊥). The σ bonding energy (ΔEσ) is the primary component of orbital interactions in all dimetallocenes. This arises from the donation of a lone pair of electrons from [CpM] to an empty spn hybrid orbital of [CpBe]+. The deformation densities Δρ1 in Fig. 5 represent this σ-bonding interaction. Hence, the σ-bonding interaction in dimetallocenes can be best described by a donor–acceptor (Be ← M) bond, with a significant contribution from electrostatic interaction (61.35%) (HOMO−2, Fig. 2, and Δρ1, Fig. 5). It is notable that the percentage of electrostatic interactions is lower than that of orbital interactions for all other heavier group 13 compounds. The donor fragment orbital of BCp is an spn hybrid orbital on B, with the back lobe strongly bonding with all in-phase π-molecular orbitals of the Cp ring (Fig. 4). As a result, there is a net charge transfer from the Cp ring to B. Since the electronegativity of heavier group 13 elements is much lower than that of B, electrostatic interactions between Cp and these heavier elements are more dominant. Consequently, the Cp ring carries a slight positive charge, while the group 13 element bears a slight negative charge in [Cp–Be–B–Cp]+. An opposite trend is observed in the heavier group 13 analogues. It is to be noted that the previously reported heterobimetallic dimetallocenes show significant ionic character for the Al–Li bond, which can be attributed to the difference in the electronegativity of Al and Li compared to Be and group-13 elements.54,55
ΔEπ∥ and ΔEπ⊥ correspond to pseudo-π bonding interactions resulting from donation of electrons from the bonding molecular orbital corresponding to Cp and Be interactions in [CpBe]+ to the antibonding molecular orbital corresponding to Cp and M interactions in [MCp]. Accordingly, these pseudo-π-interactions can be considered as hyperconjugative interactions. However, these pseudo-π-bonding interactions also show significant polarisation of electron density from the Cp ring to the M centre as well. This can be attributed to the high positive charge in the [BeCp]+ fragment. The sum of ΔEπ∥ and ΔEπ⊥ increases in the following order: Be–B (−7.98 kcal mol−1) < Be–Ga (−5.30 kcal mol−1) < Be–In (−4.86 kcal mol−1) < Be–Tl (−4.77 kcal mol−1) < Be–Al (−3.66 kcal mol−1). This can be correlated with the decrease in Be–C bond distances as the group 13 element changes from B to Tl, except for [Cp–Be–Al–Cp]+. Note that these second-order interactions are more efficient when the Be–M bond has more covalent character, which is the highest in [Cp–Be–Al–Cp]+. The deformation densities Δρ2 and Δρ3 in Fig. 5 represent the above-discussed pseudo π-bonding interaction. Since hyperconjugative interactions are of secondary type, energy stabilisation is found to be much less across all dimetallocenes. The dispersive interactions (ΔEdisp) stabilize and the extent of stabilization is similar in all dimetallocenes (−4.28 to −5.30 kcal mol−1). Hence, it is proposed that the substituted Cp rings, which could stabilize dimetallocenes more effectively through dispersive interactions, can be potential synthetic targets.
EDA considers the interactions between fragments A0 and B0 in their electronic and geometric ground states ψ0A and ψ0B with energies E0A and E0B,68 resulting in the formation of a molecule A–B with the corresponding wave function ψAB and energy EAB. The initial step involves distorting the fragments A0 and B0 from their equilibrium geometries and wave functions ψ0A and ψ0B to the electronic states and geometries ψA and ψB (with energies EA and EB) that they have in the molecule A–B. The preparation energy ΔEprep (eqn (1)) is the total energy required to electrically excite and distort all fragments to this state:
| ΔEprep = EA − E0A + EB − E0B | (1) |
The intrinsic interaction energy (ΔEint), which is the difference between the energy of the molecule EAB and the energies of the prepared fragments EA and EB (eqn (2)), is calculated in EDA:
| ΔEint = EAB − EA − EB | (2) |
The bond dissociation energy De (by definition with the opposite sign) is the sum of the instantaneous interaction energy ΔEint and ΔEprep (eqn (3))
| −De = ΔEint + ΔEprep | (3) |
The instantaneous interaction energy of a bond A–B, which is an indicator of the bond strength, is partitioned into chemically meaningful components. Within the EDA scheme, this partitioning is as follows.
| ΔEint = ΔEelect + ΔEPauli + ΔEorb + ΔEdisp |
The term ΔEelect refers to the quasi-classical electrostatic interactions, wherein the frozen electron density distribution of the fragments in the molecular geometry is computed using quantum mechanics, while the interaction between the nuclei and electrons is determined by classical rules. This is an appealing component that provides an estimate of electrostatic interactions not accounted for in the orbital-based analysis. It is important to remember that ΔEPauli denotes the repulsive interactions between fragments of a similar spin as they approach each other, not ionic bonding. It is computed using the antisymmetrization and renormalisation of the Kohn–Sham determinant of the orbitals. ΔEb is the energy associated with the stability of orbitals during bond formation and constitutes an attractive component.
It encompasses the impacts of charge transfer, relaxation, and polarisation resulting from the alteration of the charge distribution of the fragments during bond formation.
In the EDA-NOCV framework, the ΔEorb term can be further decomposed into contributions from each irreducible representation within the point group of the interacting system. Developed by Mitoraj and Michalak, the natural orbitals for chemical valence (NOCV) offer insight into each orbital interaction between the fragments and the contribution of each pair of interacting orbitals (ψ−n/ψn) to the total bond energy.
The EDA-NOCV system expresses ΔEorb in terms of NOCVs as follows:
The Kohn–Sham matrix elements
and
are defined over NOCV pairs with eigenvalues −vn and vn, respectively, with respect to the transition-state density. The density of the transition state is the intermediate density between that of the molecule and the superimposed fragments. Orbitals with a negative value exhibit antibonding properties, while those with a positive value exhibit bonding properties. It is equivalent to the percentage of electron density that is transmitted between NOCV pairs during the formation of the molecule from the frozen fragments. The strength of pairwise orbital interactions can be quantitatively estimated, and the resulting change in charge density is visualised using deformation densities.
| This journal is © The Royal Society of Chemistry 2026 |