Stability and bonding of the multiply coordinated bimetallic boron cycles: B8M22−, B7NM2 and B6C2M2 with M = Sc and Ti

Hung Tan Phamab, Nguyen Minh Tamab, My Phuong Pham-Hoc and Minh Tho Nguyen*d
aComputational Chemistry Research Group, Ton Duc Thang University, Ho Chi Minh City, Vietnam. E-mail: phamtanhung@tdt.edu.vn
bFaculty of Applied Sciences, Ton Duc Thang University, Ho Chi Minh City, Vietnam
cInstitute for Computational Science and Technology (ICST), Ho Chi Minh City, Vietnam
dDepartment of Chemistry, KU Leuven, Celestijnenlaan 200F, B-3001 Leuven, Belgium. E-mail: minh.nguyen@kuleuven.be

Received 24th February 2016 , Accepted 13th May 2016

First published on 17th May 2016


Abstract

A theoretical investigation of the geometry, stability and aromaticity of boron clusters doped by two Sc and Ti atoms was carried out using DFT calculations. The Sc and Ti atoms form the bimetallic boron cycles not only with the eight-membered ring B82− but also with the isoelectronic species B7N and B6C2 rings yielding the planar B82−, B7N and B6C2 rings sandwiched to two transition metals. The N and C atoms prefer to form eight-membered hetero-rings via the classical 2e–2c bonding rather than occupy high coordination positions. Both C atoms avoid binding each other. The thermodynamic stability of all bimetallic boron cycles is induced by the stabilizing overlap between the bonding and anti-bonding MOs of a metallic dimer M2 with levels of an eight-membered ring. These stabilizing interactions also release two sets of delocalized σ and π MOs, which obey the (4N + 2) electron count. Such a double σ and π aromaticity feature is clearly supported by the magnetic ring current flows.


1. Introduction

The diversity in geometrical structures of small boron clusters is bewildering. While it is well known that the B12 icosahedral three-dimensional (3D) cage is a structural unit in many boron polyforms,1 this structure is found to be much less stable than its planar isomer.2 Previous studies on a large number of pure boron clusters show that the bare boron clusters basically have four different classes of structures involving the planar (or quasi-planar), the tubular (double ring, triple ring and multiple ring…), the bowl shaped, and the fullerene-like and cage forms. Most of the boron clusters Bn having sizes smaller than n < 20 atoms are planar or quasi-planar,3 whereas the lowest-lying structures of the even sizes B2m in the range of 2m = 20–26 have a double ring form. The latter is the simplest tubular structure, which consists of two Bm strings often connected in an anti-prism fashion.4 The triple ring form is observed in B27 and B42 which result from a superposition of three B9 and B14 strings in an anti-prism motif, respectively.5,6 Regarding the cationic state, 3D structure is already found for the size of B17+ (ref. 7) whereas the planar and quasi-planar forms remain the global minimum structures for anionic boron clusters up to the size of B27.5,8 The B30, B32 and B36 clusters present the bowl shape which contains a large hole, and their electron distribution satisfies the disk-aromaticity.6,9,10 A fullerene motif of structure is established in the case of B38 and B40 clusters.11–13 Recently, the B44 cluster has been shown to have a cage-like structure containing two nonagonal holes. Such a feature is unprecedented in elemental clusters.14

Another class of boron clusters, represented firstly by the anions B82− and B9, has planar cyclic geometry in which a B atom occupies the central position of the B7 and B8 rings, respectively.15 It is remarkable that a doping of one transition metal M gives rise to the M@Bn planar cyclic structures in which the M atom is surrounded by the B8, B9 and B10 rings.8 The stability of the planar metallic boron cycles can be rationalized in terms of strong orbital interactions where the transition metal uses its d-AOs to form two delocalized bonding patterns. The latter satisfy the classical (4N + 2) electrons count in both series of delocalized π and σ MOs therefore they are classified as doubly π and σ aromatic. Stimulated by the finding that the B@B5H5+ cluster is a planar cycle,16 in which one B atom occupies the central position of the B5H5+ pentagonal cycle, a number of doped M@BnHn cyclic forms have been established.17 The planarity of these systems are rationalized by the aromaticity with 10 π electrons. Although several boron rings centered by other main group elements were previously investigated, none of them was found to be a global minimum on the corresponding potential energy surface.18,19 The central atom needs to form delocalized π and σ bonds with the Bn ring, but when a main group element having a larger electronegativity than B, it favors connection to other atoms through simpler 2 center–2 electrons (2c–2e) bonds. As for typical examples, carbon–boron mixed cluster reveals many B@BxCy boron hetero-cycles where C atoms participate to ring not in central position.20–23 The electronic partition analysis shows that C atoms provide to the delocalized π and σ MOs and subsequently stabilize boron cycles.

For the boron-based clusters multiply doped by transition metal atoms, some organometallic compounds were observed. In the series of bimetallic complexes BnXnRe2, the boron ring is uniformly coordinated with two Re atoms. Of the derivatives B5X5 and B6X6, the boron units were found to be perfectly planar pentagon and hexagon motifs, respectively.24,25 The planar B6 ring was found to play as a key structural motif in the solid compound Ti7Rh4Ir2B8,26 in which the B6 unit is sandwiched by two Ti atoms. Recently, we reported on a novel motif of structure in which two metals are coordinated vertically and oppositely with the planar boron ring.18 In these bimetallic boron cycles having Fe and Co as dopants, the global energy minima of the resulting B7Co2, B7Fe2 and B7CoFe clusters contain each a perfectly planar B7 cycle, which is sandwiched with two metallic atoms vertically placed along the C7 axis.27 The high thermodynamic stability of these bimetallic boron cycles arises from a combination of two stabilizing effects. On the one hand, the M2 dimeric metals introduce their electrons to fill up the empty levels of the B7 string, and on the other hand, both anti-bonding and bonding MOs of M2 are significantly stabilized through orbital interactions with the cyclic B7 counterpart.

Although bimetallic boron clusters have been identified in the seven-membered boron rings with Fe and Co as dopants, the derivatives containing larger boron rings such as eight-membered ring are not known yet. As mentioned above, the main group elements such as the C and N atoms tend to prefer formation of classical 2c–2e bonds, in part due to their higher electronegativity. As a prediction, these atoms possibly connect with boron atoms to reveal hetero-cycles and subsequently, in the presentation of metals, the bimetallic hetero-boron cycles could be established. In this context, we set out to search for new members of the family of bimetallic boron cycles containing eight-membered rings using quantum chemical computations. Our extensive theoretical search on the geometry and electronic structure of several B8M2 species reveals that the M = Sc and Ti atoms are quite suitable for this purpose. For the eight-membered rings, beside the neutral B8 and its dianionic B82− ring, simple substitutions generating the B7N ring (with substitution of B2− by N) and the B6C2 ring (with replacement of two B anions by two C atoms) are also considered. The latter systems allow us to analyze the behavior of the series of isoelectronic systems. The electronic structure and chemical bonding feature of this class of bimetallic boron cycles will be examined using the electron localization indicator. Their aromatic character is further probed using the magnetic response of the electron density as described by the ring current maps.

2. Computational methods

The potential energy surfaces of the B8M2, B8M22−, B7NM2 and B6C2M2 clusters, with M = Sc and Ti, are explored using the stochastic genetic algorithms.28 In addition, the guess structures are also manually built by adding two metal atoms to the framework of the pure Bn species. Initial B7NM2 and B6C2M2 structures obtained following substitution of a single B atom by a N, or two B atoms by two C are also generated on the basis of the B8M22− counterparts.

Geometric optimizations and harmonic vibrational analyses are carried out using density functional theory (DFT) with the hybrid TPSSh,29 which was previously shown to give better results on relative energies for boron clusters than those obtained with other available functionals, as compared to the high accuracy MO coupled-cluster method (CCSD(T)).30 The PBE1PBE31 and BPW91 (ref. 32) functionals are found to excellently predict geometry and spin state of Fe(C6H6)n systems, therefore these functional are use in the current investigation where the B6C2 and B7N containing structures are considered.33 In order to evaluate the reliability of three functionals, the bond length, vibrational frequency, dissociation energy and electron affinity (EA) of the dimers Sc2 and Ti2 are calculated and compared to available experimental values.34–40 As shown in Table 1, in comparison to available experimental values, the TPSSh and PBE1PBE in combining with the cc-pVTZ basis set produce better results than the BPW91/cc-pVTZ calculation.

Table 1 Bond length (d(M2), angstrom), vibrational frequency (Freq., cm−1), dissociation energy (Do, eV) and electron affinity (EA, eV) of Sc2 and Ti2 obtained by TPSSh, PBE1PBE and BPW91 calculations and compared to experiment (Exptl)
  Exptl TPSSh PBE1PBE BPW91
d(M2) Freq. Do d(M2) Freq. EA Do d(M2) Freq. EA Do d(M2) Freq. EA Do
a CASSCF calculations taken from ref. 33.b Ref. 36 and 37.c Ref. 35.d Ref. 38.e Ref. 39.f Ref. 40.
Sc2 2.51a 238.9c 1.65e 2.6 258.3 0.68 1.1 2.6 267.3 0.94 0.76 2.6 241.3 2.9 −0.5
Ti2 1.94b 407.9d 1.54f 1.90 502.3 0.35 2.0 1.95 536.8 1.0 1.2 1.90 470.7 0.7 2.2


The initial structures are first optimized using the small 3-21G(d) basis set.41,42 The optimized lower-energy isomers, being within a range of 50 kcal mol−1 on relative energies, are subsequently reoptimized using the same TPSSh functional, but in conjunction with the larger cc-pVTZ basis set.43,44 Electronic structure theory computations are performed using the Gaussian 09 suite of program.45

The magnetic response is calculated using the CTOCD-DZ method43 implemented in the SYSMO program,46,47 which is connected to the GAMESS-UK package.48 The ring current approach49,50 is calculated using the B3LYP/6-311G(d) level. In the current density map, the contour and shading show the modulus of induced current density and narrows display its projection on plotting plane. As for a convention, anticlockwise or clockwise circulations correspond to diatropic or paratropic current, respectively. The diatropic current flow indicates an aromatic character, whereas the paratropic current flow suggests an anti-aromatic character.

3. Results and discussion

3.1 Geometries of the B8M2, B8M22−, B7NM2 and B6C2M2 clusters

Only the eight-membered boron cycles are considered in this investigation. As for a convention, the structure is noted as M2.X.n, in which M2 = Sc2 and Ti2 stands for two metal dopants, X = B8, C2 and N according to the pure B8 cycle and the substituted derivatives by two and one B atoms by two C and one N atoms, respectively, and finally n = 1, 2, 3… denotes the isomers with increasing energy ordering. When relative energies are mentioned, the values are consistently referred to the corresponding most stable isomer n = 1. In the dianionic B8M22−, the isomers are referred to as M2.B8.da.n.

Fig. 1 displays the shape of the energetically lower-lying isomers of both series B8M2 and B8M22− with M = Sc and Ti obtained using the TPSSh/cc-pVTZ level. Geometry of B6C2M2 and B7NM2 are displayed in Fig. 2. As for a comparison, their relative energy obtained by using three different TPSSh, PBE1PBE and BPW91 functionals in various spin states are listed in Table 1.


image file: c6ra04948f-f1.tif
Fig. 1 Energetically lower-lying isomers of (a) B8Sc2, (b) B8Sc22−, (c) B8Ti2 and (d) B8Ti22− clusters. Geometry optimizations and relative energies (kcal mol−1 with ZPE corrections) were obtained using the TPSSh/cc-pVTZ computations.

image file: c6ra04948f-f2.tif
Fig. 2 Lower-lying isomers of (a) B6C2Sc2 and B6C2Ti2 clusters, and (b) B7NSc2 and B7NTi2 clusters. Geometry optimizations and energies (kcal mol−1 with ZPE corrections) were performed using the TPSSh/cc-pVTZ computations.

For the neutral B8Sc2 cluster, the triplet spin state Sc2.B8.1 is confirmed as the ground state by TPSSh, PBE1PBE and BPW91 calculations whereas the quintet multiplicity is highly unstable. The structural characteristic of Sc2.B8.1 shows that two Sc atoms are vertically coordinated to the planar B8 string. All DFT calculations demonstrate that the next isomer Sc2.B8.2, in three spin states, is extremely unstable. In the following section, the structural motif in which two metal atoms are vertically sandwiched in the opposite sides of a planar eight-membered ring is referred to as a ‘bimetallic planar cyclic cluster’. Following attachment of two electrons, the bimetallic cyclic structure Sc2.B8.da.1 is also found to be the lowest energy isomer. TPSSh/cc-pVTZ calculations point out that the quintet multiplicity is the ground state of Sc2.B8.da.1, whereas both PBE1PBE/cc-pVTZ and BPW91/cc-pVTZ computations predict the triplet spin state as the ground state. The fish-like structure Sc2.B8.da.2 is predicted to be highly unstable by all DFT calculations (Table 1).

The boron cyclic form is not favored in the neutral B8Ti2. The fish-like structure Ti2.B8.1 (Cs 1A′) in which both metal atoms are coordinated to a B6 string and two B atoms capped at two different sites of the same B–B edge, is the ground state of B8Ti2. The Ti2.B8.2 bimetallic cycle (D8h 1A1g) is much less stable as indicated by TPSSh, PBE1PBE and BPW91 calculations (Table 1). However, the bimetallic boron cycle Ti2.B8.da.1 turns out to be again the lowest-energy structure for the dianion B8Ti22−, being ∼10 kcal mol−1 below the corresponding fish-like structure (Table 1). It should be stressed that both Sc and Ti dopants give rise to the bimetallic cyclic boron structures with high symmetry in the dianion state.

Similar to the B8Ti22− and B8Sc22− dianions, both isoelectronic neutral B6C2Sc2 and B6C2Ti2, where the two B centers of B8 are replaced by two C atoms, remain stable in the bimetallic cyclic motif with the emergence of a B6C2 planar heterocycle. In the case of B6C2Sc2, the TPSSh, PBE1PBE and BPW91 calculations show that the singlet Sc2.C2.1 and Sc2.C2.2 isomers are degenerate with an energy gap of <1 kcal mol−1, and as a consequence they can be regarded as competing ground state for the neutral mixed B6C2Sc2. Within these structures, two Sc atoms are coordinated vertically but oppositely to the B6C2 perfect planar ring.

Similar structure is observed also in the case of Ti-dopant. As shown in Fig. 2a, a hetero-boron cycle structure, Ti2.C2.1, is identified as the ground state of B6C2Ti2 with high symmetry and low spin multiplicity (singlet state) by all functionals used. Some other bimetallic boron cycles are observed on the potential energy surface of B6C2Ti2, but they are significantly less stable than Ti2.C2.1. It is remarkable that the bimetallic boron cycles B6C2M2 with M = Sc and Ti containing C–C connection are highly unstable, which results from the fact that in the formation of bimetallic carbon–boron cycles with two C atoms, the C dopants avoid to meet and to bind to each other.

Substitution of a single B site of B8Sc22− and B8Ti22− by an N atom also releases a bimetallic cyclic structure, Sc2.N.1 and Ti2.N.1 again with a planar B7N hetero-ring. The fish-like structures Sc2.N.2 and Ti2.N.2 are found at ∼10–14 kcal mol−1 above the corresponding ground state, predicted by TPSSh, PBE1PBE and BPW91 calculations, respectively. The next isomers Sc2.N.3 and Ti2.N.3 are highly unstable in both singlet, triplet and quintet states.

Overall, the dianion B8M22− with M = Sc and Ti and their isoelectronic neutral B6C2M2 and B7NM2 clusters provide us with the new members of the class of bimetallic boron cycles. Of these species, it is clear that Sc2.C2.1, Sc2.N.1, Ti2.C2.1 and Ti2.N.1 present the bimetallic hetero-boron cycles in which two metals are coordinated to B6C2 and B7N hetero-ring. The most interesting result here is that both N and C atoms participate together with boron atoms in the formation of the eight-membered hetero-ring rather than to be located in highly coordinated position as the two metals.

3.2 Energetic properties

In order to explore the stability of bimetallic boron cycles, the dissociation energy (DE) where parent structure dissociates giving a metallic dimer plus the corresponding ring, is calculated using three functionals and given in Table 3. Additionally, the binding energy (BE, Table 4) are also performed to quantify their stability. The DE values of the channel in which B8M2 and B8M22− structures release the global minimum B8 and B82− plus M2 dimers, are significantly lower (Table 2) and illustrates that B8M2 and B8M22− favor the dissociation channel giving B@B7 and B@B72−. The DE value of the B8M22− dianion is somewhat lower than that of the neutral. The connection between M2 and B82− dianionic ring becomes thus weaker than that between M2 and B8 neutral ring. This is supported by binding energy calculations in which BE value of neutral B8M2 is greater than the corresponding B8M22−. Similarly, in comparison to B8M2 neutral, DE values of B7NM2 and B6C2M2 cluster with M = Ti and Sc are smaller, therefore the interactions between M2 and B7N and B6C2 rings are weaker than that of in the B8M2 neutral. Although the substitution of B atom by N and C atoms in B82− ring results in many new BN, CB bonds which are stronger than the BB bond in B82− ring, the interactions of two M atoms with the rings either in B7NM2 and B6C2M2 or in B8M2 and B8M22− cycles are of the same order of magnitude.
Table 2 The relative energy (kcal mol−1) of B8M2, B8M22−, B6C2M2 and B7NM2 clusters with M = Sc and Ti at various spins states using TPSSh, PBE1PBE and BPW91 functionals combining with cc-pVTZ basis set
Structure TPSSh PBE1PBE BPW91
Singlet Triplet Quintet Singlet Triplet Quintet Singlet Triplet Quintet
Sc2.B8.1 3.4 0.0 48.7 4.0 0.0 48.0 4.0 0.0 44.3
Sc2.B8.2 22.9 47.5 68.7 22.3 46.0 66.3 22.2 38.1 60.6
Ti2.B8.1 0.0 15.5 41.8 0.0 21.8 41.8 0.0 15.4 42.0
Ti2.B8.2 6.1 6.4 11.8 10.1 9.5 11.4 5.7 7.7 15.9
Ti2.B8.3 14.9 18.8 42.6 15.0 17.6 38.3 12.5 17.7 43.2
Sc2.B8.da.1 6.8 15.2 0.0 12.3 0.0 16.8 15.0 0.0 4.2
Sc2.B8.da.2 25.1 20.3 31.3 38.7 27.3 33.1 33.4 26.1 33.6
Ti2.B8.da.1 0.0 3.5 0.5 0.5 0.0 0.0 0.0 15.9 6.9
Ti2.B8.da.2 10.3 9.0 17.8 8.3 1.1 27.0 9.4 6.7 18.5
Ti2.B8.da.3 28.9 20.4 23.4 26.7 16.2 22.4 27.0 19.5 23.4
Sc2.C2.1 0.0 15.7 49.7 0.0 16.0 45.7 0.0 14.5 47.3
Sc2.C2.2 0.6 20.2 45.2 0.5 20.6 45.7 0.7 18.7 43.4
Sc2.C2.3 6.5 22.6 47.0 6.6 22.3 47.5 6.1 19.5 45.3
Ti2.C2.1 0.0 6.8 11.0 0.0 4.0 6.0 0.0 1.5 14.1
Ti2.C2.2 6.6 4.2 17.5 6.8 1.5 12.6 6.2 6.0 16.5
Ti2.C2.3 12.1 6.1 16.4 12.0 12.1 11.7 10.2 7.4 18.8
Sc2.N.1 0.0 10.1 34.3 0.0 8.2 35.6 0.0 8.9 32.0
Sc2.N.2 13.1 18.4 46.7 11.3 17.7 36.3 11.9 17.1 35.7
Sc2.N.3 22.5 41.0 71.9 20.8 39.4 74.3 21.1 40.3 68.1
Ti2.N.1 0.0 10.6 19.0 0.0 14.3 14.1 0.0 15.1 26.2
Ti2.N.2 11.6 14.6 38.6 10.5 10.7 27.8 14.6 17.5 41.7
Ti2.N.3 18.7 40.0 43.1 17.8 34.7 36.5 21.7 41.1 51.0


Table 3 The dissociation energy of bimetallic cycles obtained by TPSSh, BPW91 and PBE1PBE calculations
  TPSSh BPW91 BPE1PBE
B8Sc2 → B@B7 + Sc2 230.8 289.7 224.6
B8Sc22− → B@B72− + Sc2 174.8 216.0 171.5
B8Ti2 → B@B7 + Ti2 218.9 306.9 215.2
B8Ti22− → B@B72− + Ti2 165.0 232.9 156.9
B8Sc2 → B8 + Sc2 328.0 312.0 333.5
B8Sc22− → B82− + Sc2 332.1 271.2 288.3
B8Ti2 → B8 + Ti2 327.4 406.6 334.9
B8Ti22− → B82− + Ti2 286.5 299.7 289.8
B7NSc2 → B7N + Sc2 286.8 271.1 289.0
B7NTi2 → B7N + Ti2 259.9 257.9 261.1
B6C2Sc2 → B6C2 + Sc2 305.5 289.7 309.5
B6C2Ti2 → B6C2 + Ti2 279.4 272.5 282.2


Table 4 The binding energy (D0, kcal mol−1) of bimetallic cycles obtained. Using three different functionals with the cc-pVTZ basis set
  TPSSh BPW91 PBE1PBE
B8Sc2 106.8 110.1 109.3
B8Sc22− 104.8 108.8 107.8
B8Ti2 107.7 111.4 109.3
B8Ti22− 105.9 110.0 107.3
B7NSc2 115.5 114.9 118.0
B7NTi2 117.0 117.0 118.3
B6C2Ti2 126.0 130.0 127.6
B6C2Sc2 125.5 129.2 128.4


It should be noted that although the B8Ti2 bimetallic cycle is not the global minimum, this structure has a greater DE value than B8Ti22− cycle which is determined as the ground state. In combination with calculations of the binding energy, these results show that the interaction of M2 dimer with B8, B82−, B7N and B6C2 rings which is one main contributor to the stability of bimetallic boron cycles, is quantitatively comparable. The delocalization or aromaticity is more important in stability, particular for di-anions B8M22−, B7NM2 and B6C2M2 cycles, as shown in following sections.

3.3 Metal–metal bonding

For an understanding on the behavior of the two metallic atoms, the Wiberg bond indexes (WBI)51,52 are performed for both free metal dimers, and the M–M bonds in bimetallic boron cycles, and the calculated results are given in Table 5. The bond lengths of the free diatomic Sc2 and Ti2 amount to 2.8 and 1.94 Å,53 respectively, and WBI values are found as 0.25 and 0.78. As seen from Table 5, the bond lengths calculated for the M2 connection in all bimetallic boron cycles including B8M22−, B6C2M and B7NM2 is ∼0.5 Å longer than those of free metal dimers, while the WBI values of M–M connections change according to the nature of the dopant. The B8M22− clusters have WBI values of ∼0.6 for Ti2 and Sc2, but these are of 0.9 and 0.5 for B7NTi2 and B7NSc2 clusters, respectively. For B6C2M2, WBI values amount to 1.2 and 0.5 for Ti and Sc dopants, respectively. On the geometrical aspect, formation of bimetallic boron cycles tends to increase distance between two metal atoms. Although the WBI value of M2 in B8M22− is reduced, the B6C2M2 and B7NM2 result in larger values of WBI for the M2 unit. According to the meaning of WBI, the presence of C or N dopants tends to enhance the strength of M–M bonds, in spite of their longer distances.
Table 5 Bond length of M–M, B–M, C–M, N–M, C–B and N–B (Å) and WBI values of M–M connections of the bimetallic boron cycles
Species Parameter Ti Sc
a Value obtained for B8M22− cyclic structure.
M2 d(Exptl) 1.94 2.51
d(M2) 1.90 2.6
WBI 0.78 0.25
B8M2/B8M22− d(M–M) 2.92, 2.49 2.94, 2.83
d(B–M)a 2.4 2.4
WBI 0.64 0.65
B7NM2 d(M–M) 2.7 2.9
d(B–M) 2.3 2.4
d(N–M) 3.0 2.3
d(N–B) 1.4 1.4
WBIM–M 0.9 0.49
B6C2M2 d(M–M) 2.5 2.9
d(B–M) 2.3 2.4
d(C–M) 2.4 2.3
d(C–B) 1.4 1.4
WBIM–M 1.2 0.48


3.4 Orbital interactions

To probe further the chemical bonding feature and thereby the stability of bimetallic boron cycles, it is imperative to consider the bonding model which raises from d–d interactions.54 As for a representative example, Fig. 3 illustrates the orbital interactions of the Ti2 dimer with the B82− string. The results obtained for Ti2 can be generalized for others bimetal-doped boron cycles.
image file: c6ra04948f-f3.tif
Fig. 3 Orbital interaction diagram of the Ni2 dimer with the B82− string.

The Ti atom has an electron configuration of [Ar](3d)2(4s)2, and as a consequence, the Ti2 dimer has enough electrons to fully occupy an electronic configuration of [σ24sσ23dπ4]. The two σ MOs include the σ4s orbital which is resulted from overlap of the 4s AOs, and the σ3d which is a combination of 3d AOs. The anti-bonding δ*, π* and σ* MOs are filled, and thereby, the strength of metal–metal bond is reduced.

The delocalized MO pattern of the ring B82− appears to satisfy a disk aromatic framework, which arises from a model of a particle in a circular box.55 Within this model, the boundary condition gives two quantum numbers, namely the radial quantum number n and the rotational quantum number m. The radial quantum number has values of n = 1, 2, 3, … whereas m = 0, ±1, ±2, … which is denoted as m = σ, π, δ, …, respectively. The state with non-zero value of m shall thus be doubly degenerate. The lowest-lying eigenstates in ascending energy are 1σ, 1π, 1δ… etc. The full occupation of degenerate eigenstates which correspond to 2, 6, 12, 16, 20, … electrons leads to a disk aromatic character.

The orbital interaction of the B82− disk configuration and the Ti2 bond inherently enhances the stability of boron cycles, as displayed in Fig. 3. From there, the π*-MOs of the Ti2 dimer whose occupation significantly reduces the strength of the Ti–Ti bond, are stabilized upon interaction with the doubly degenerate 1π levels of either the ring B7 or B8, and thereby produces the π levels of the entire system. Interestingly, the δ*-MOs of Ti2 enjoy a stabilizing overlap with the vacant 1δ level, and releases the doubly degenerate δ* levels, as given in Fig. 3. Interaction of the bonding σ4s and the 2σ level ends up in a creation of the σ4s MOs for the B8Ti22− bimetallic dianion. The π bonding MOs of Ti2 have an enhanced overlap with the 2π level of the B82− disk, whereas the 2δ-MOs of the B82− counterpart undergo a stabilizing interaction with the δ-MOs of Ti2 yielding the δ levels of the complex. As a result, the orbital configuration of the B8Ti22− cyclic structure as […(σ4s)2(π)4(π*)4(δ)2(δ*)4] is now fully occupied by 16 electrons.

The electron configurations of the different bimetallic cyclic boron clusters are listed in Table 6. The π* and δ* levels of the isoelectronic cycles to B8Ti22− involving B6C2Ti2 and B7NTi2 are fully occupied (Table 6). As a result, they tend to enhance the stability of cyclic structures, although they reduce the strength of Ti–Ti connection. The σ4s and π levels are actually fully occupied, whereas the δ molecular orbitals are filled by two (2) electrons. These MOs increase the strength of Ti–Ti bond, and stabilize consequently the bimetallic cyclic boron clusters. The dianion B8Sc22− has the orbital configuration of …(σ4s)2(π*)4(π)4(δ*)4 in which the π* and δ* MOs are fully occupied, as given in Table 2. The π* and δ* MOs of the isoelectronic boron cycles with B8Sc22− involving B7NSc2 and B6C2Sc2 are also fulfilled by 4 electrons for each cluster. It is obvious that they are significant contributor to the stability of boron cycles containing the Sc-dopant. The full occupation of σ4s and π levels contributes significantly to the thermodynamic stability of cyclic structure, due to the enhanced Sc–Sc connection. Therefore, a successive occupation of σ4s, π, π* and δ* levels leads to the formation of bimetallic boron cycles. It is clear from Table 6 that the B6C2Ti2 and B7NTi2 have different MO patterns, in which the appearance of C and N atoms, that are more electronegative than B atom, leads to a substantial change in the MO energy levels with expected splitting of the doubly degenerate MOs.

Table 6 Orbital configurations of the bimetallic boron cycles considered
Cluster Orbital configuration
B8Sc22− 4s)2(π*)4(π)4(δ*)4
B8Ti22− 4s)2(π*)4(π)4(δ*)4(δ)2
B7NSc2 4s)2(π*)2(π)2(π*)2(π)2(δ*)4
B7NTi2 (π*)24s)2(π*)2(π)4(δ*)43d)2
B6C2Sc2 4s)2(π)2(π*)2(π)2(π*)2(δ*)43d)0
B6C2Ti2 4s)2(π)2(π*)2(π)2(π*)2(δ*)43d)2


Fig. 4 displays a correlation diagram of three isoelectronic bimetallic boron cycles B8Ti22−, B6C2Ti2 and B7NTi2. As expected, the doubly degenerate levels of the B8Ti22− cycle including the π, π* and δ* are now split into two separate levels in both B6C2Ti2 and B7NTi2 cycles. Due to the influence of the more electronegative C and N atoms, the energy of the σ4s, π*, π, δ* and δ levels of B6C2Ti2 and B7NTi2 become higher than those of the B8Ti22−. Overall, the orbital interactions demonstrate that the high thermodynamic stability of bimetallic boron cycles is a consequence of two factors, namely, (i) the empty levels of B82− involving 2π, 1δ and 2δ are occupied, and (ii) the bonding and anti-bonding MOs of M2 are involved in stabilizing interactions with the levels of the disk B82−, which invariably enhance the stability of the resulting clusters. As a consequence, the B8M20/2−, B7NM2 and B6C2M2 clusters are stabilized within the bimetallic cyclic motif.


image file: c6ra04948f-f4.tif
Fig. 4 Orbital correlation diagram for the B8Ti22−, B6C2Ti2 and B7NTi2 clusters.

3.5 An analysis of the electron distribution

In order to explore further the chemical bonding phenomena, an analysis of the electron density using the electron localization indicator (ELI-D) technique56 is performed, and the resulting partition of the densities is shown in Fig. 5. A similar localization domain pattern is thus observed for all bimetallic clusters considered. The di-synaptic basin V(B,B) is found for two B atoms of the B8 ring in both B8Sc22− and B8Ti22−, therefore the B atoms connect with each other via the classical 2e–2c bonds, and the latter go around the B8 planar ring. For B7NM2, it has two localization domains located between N dopant and two B atoms yielding two 2e–2c bonds of the N with B atoms. The same behavior is observed in the mixed carbon–boron B6C2M2 in which the C dopants connect with other B atoms by formation of 2e–2c bonds, as reflected by the presence of the V(C,B) basins. No localization domain, located in the regions between the metal M and the planar rings involving B8, B7N and B6C2, is observed, indicating a certain ionic character of the M–X bond with X = B, C and N. The localization centered on the M–M bonding is observed only at lower bifurcation values, due to their large radii. However, this does not imply that there is no direct M–M bonding.
image file: c6ra04948f-f5.tif
Fig. 5 The ELI_D isosurface maps obtained at the bifurcation value of 1.4 for the B8M22−, B6C2M2 and B7NM2 bimetallic cyclic boron clusters, based on densities obtained at the TPSSh/cc-pVTZ level.

3.6 The aromaticity

The aromatic property of the bimetallic boron cycles considered is now analyzed using the ring current concept, which arises from a magnetic response of the electron density of the whole system. The magnetic response can be partitioned in terms of MO contributions. The contribution of each MO to the total ring current can be calculated, and thereby the participation of each MO to the aromatic character of the whole cluster can be established. The current density maps of both π and σ systems of electrons are produced on the B8, B6C2 and B7N molecular planes. Fig. 6 displays the ring current flows of the π, σ and total (involving π and σ electrons) for the bimetallic boron cycles, in their molecular plane. The ring current density maps of the π and σ orbitals are illustrated in Fig. 7.
image file: c6ra04948f-f6.tif
Fig. 6 The total, π and σ ring current maps of isoelectronic systems B8M22−, B7NM2 and B6C2M2 clusters with M = Sc, Ti (B3LYP/6-311G(d)).

image file: c6ra04948f-f7.tif
Fig. 7 The ring current maps of π-MOs of B8Ti22−, B7NTi2 and B6C2Ti2 clusters (B3LYP/6-311G(d)).

All boron cycles containing the Sc-dopant, including B8Sc22−, B6C2Sc2 and B7NSc2, exhibit a diatropic current density for both π and σ systems of electrons (Fig. 6a). A similar character is identified for both B8Ti22− and B7NTi2 structures whose diatropic ring current maps are observed for both π and σ electrons (Fig. 6b). Therefore they can be classified as having a double π and σ aromaticity, according to the magnetic criteria. For the neutral B6C2Ti2 cluster, the ring current of π electrons is of diatropic nature in magnetic response, whereas its σ electrons produce a complicated ring current map. Therefore, the aromatic character of B6C2Ti2 will be considered in a following section on the basis of MO contributions.

It is important to consider the contribution of each MO to ring current maps, in such a way that its participation to the aromaticity can be revealed. Each of the bimetallic boron cycles B8M22−, B7NM2 and B6C2M2 has 10 π electrons that populate the π* and δ* levels, and the MO-34. As a result, they can be classified as aromatic species, consistent with the classical (4N + 2) Hückel electron count. For a more complete understanding, the ring current maps of their π-MOs are calculated and displayed in Fig. 7 for boron cycles of Ti metal, and in Fig. S1 of the ESI file for cycles containing Sc dopants. It is clear that the π ring current flows are mainly contributed, on the one hand, by two doubly degenerate MOs, namely the δ* and π* orbitals. On the other hand, the overlap of anti-bonding δ* and π*-MOs of the dimeric metal with δ and π levels of the disk configuration achieves the π diatropic current for bimetallic boron cycles. For σ electrons, the doubly degenerate π-MO brings in the main participation to diatropic ring current maps (Fig. 8), except for the B8Ti22−. In Sc-containing boron cycles, the six σ electrons of boron cycles populating the σ4s and π levels are again consistent with indication of ring current criterion. Subsequently, the (4N + 2) electron count also works for σ electrons of Sc clusters. Similar results are obtained for the B6C2Ti2 and B7NTi planar cycles in which the π levels give the main contribution to the ring current. The σ3d MO of B6C2Ti2 is active in the magnetic response and produces a complicated and unclear map for σ electrons. However, this MO can be ignored in the consideration of σ aromaticity due to the fact that it is mainly contributed by σ3d of Ti2 dimer, rather than a delocalized MO.


image file: c6ra04948f-f8.tif
Fig. 8 The ring current maps of σ-MOs of B8Ti22−, B7NTi2 and B6C2Ti2 clusters (B3LYP/6-311G(d)).

In of the B8Ti22− cycle, there are 8 σ electrons arising from contribution of the doubly occupied δ-MO, beside the σ4s and π levels, in such a way that this species can be regarded as antiaromatic according to the (4N + 2) electron count. The ring current map established for σ electrons of B8Ti22− clearly illustrates the diatropic nature, and as a consequence, the B8Ti22− cycle should be an aromatic species. On the other hand, the (4N + 2) rule does not work in the case of B8Ti22−, simply because the δ-MO does not contribute to the ring current. Overall, the combination of the π-MOs (M2) with π levels of the eight-membered ring does not only enhance the stability of the resulting MOs, but also induce an aromatic character of the bimetallic cyclic boron clusters.

4. Concluding remarks

In the present theoretical study, we investigated on the geometry, stability and aromaticity of a series of bimetal-doped boron clusters. The Sc and Ti dopants establish the bimetallic cyclic form with the eight-membered boron ring B82− and their isoelectronic derivatives B7N and B6C2, in which a perfect planar B82−, B7N and B6C2 hetero-rings are coordinated with two transition metal atoms.

The thermodynamic stability of these bimetallic boron cycles, that are the global energy minimum of the corresponding systems, can be understood as the result of a stabilizing overlap between bonding and anti-bonding MOs of M2 with different levels of eight-membered ring. The C and N elements, which are more electronegative than the B atom, also enjoy formation of planar eight-membered ring, via the classical 2c–2e bonding, rather than occupy a high coordination position. The double aromaticity feature, which comprises both σ and π aromaticity, is clearly supported by the magnetic responses of the electron densities within the planar cycles. Nevertheless, such an aromatic character is not always in line with the classical electron count for both delocalized electron systems, and this suggests that the latter cannot reliably be applied to predict the aromatic character of this class of compounds.

Acknowledgements

We thank Ton Duc Thang University (TDTU-DEMASTED) for support. MTN is indebted to the KU Leuven Research Council (GOA program) and FWO-Vlaanderen for continuing support. The authors are grateful to Dr Remco Havenith at University of Groningen, for assistance with the computations of ring current maps.

References

  1. B. Albert and H. Hillebrecht, Angew. Chem., Int. Ed., 2009, 48, 8640 CrossRef CAS PubMed.
  2. B. T. Tai, D. J. Grant, M. T. Nguyen and D. A. Dixon, J. Phys. Chem. A, 2010, 114, 994 CrossRef PubMed.
  3. B. T. Tai, N. M. Tam and M. T. Nguyen, Chem. Phys. Lett., 2012, 530, 71 CrossRef , http://www.sciencedirect.com/science/article/pii/S0009261412000863.
  4. (a) B. T. Tai and M. T. Nguyen, Phys. Chem. Chem. Phys., 2015, 17, 13672 RSC; (b) H. T. Pham, L. V. Duong and M. T. Nguyen, J. Phys. Chem. C, 2014, 118, 24181 CrossRef CAS.
  5. L. V. Duong, H. T. Pham, N. M. Tam and M. T. Nguyen, Phys. Chem. Chem. Phys., 2014, 16, 19470 RSC.
  6. H. T. Pham, L. V. Duong, N. M. Tam, M. P. Pham-Ho and M. T. Nguyen, Chem. Phys. Lett., 2014, 608, 295 CrossRef CAS.
  7. B. T. Tai, N. M. Tam and M. T. Nguyen, Theor. Chem. Acc., 2012, 131, 1241 CrossRef.
  8. A. P. Sergeeva, I. A. Popov, Z. A. Piazza, W. L. Li, C. Romanescu, L. S. Wang and A. I. Boldyrev, Acc. Chem. Res., 2014, 47, 1349 CrossRef CAS PubMed.
  9. T. B. Tai and M. T. Nguyen, Chem. Commun., 2015, 51, 7677 RSC.
  10. Z. A. Piazza, H. S. Hu, W. L. Li, Y. F. Zhao, J. Li and L. S. Wang, Nat. Commun., 2013, 5, 3113 Search PubMed.
  11. J. Lv, Y. Wang, L. Zhu and Y. Ma, Nanoscale, 2014, 6, 11692 RSC.
  12. T. B. Tai and M. T. Nguyen, Nanoscale, 2015, 7, 331 Search PubMed.
  13. H. J. Zhai, Y. F. Zhao, W. L. Li, Q. Chen, H. Bai, H. S. Hu, Z. A. Piazza, W. J. Tian, H. G. Lu, Y. B. Wu, Y. W. Mu, G. F. Wei, Z. P. Liu, J. Li, S. D. Li and L.-S. Wang, Nat. Chem., 2014, 6, 732 Search PubMed.
  14. T. B. Tai and M. T. Nguyen, Chem. Commun., 2016, 52, 1653 RSC.
  15. H. J. Zhai, A. N. Alexandrova, K. A. Birch, A. I. Boldyrev and L. S. Wang, Angew. Chem., Int. Ed., 2003, 42, 6004 CrossRef CAS PubMed.
  16. H. L. Yu, R. L. Sang and Y. Y. Wu, J. Phys. Chem. A, 2009, 113, 3382 CrossRef CAS PubMed.
  17. J. Hou, Q. Duan, J. Quin, X. Shen, J. Zhao, Q. Liang, D. Jiang and S. Gao, Phys. Chem. Chem. Phys., 2015, 17, 9644 RSC.
  18. Z. Pu, K. Ito, P. v. R. Schleyer and Q. S. Li, Inorg. Chem., 2009, 48, 10679 CrossRef CAS PubMed.
  19. K. Ito, Z. Pu, Q. S. Li, P. v. R. Schleyer and Q. S. Li, Inorg. Chem., 2008, 47, 10906 CrossRef CAS PubMed.
  20. L. M. Wang, W. Huang, B. B. Averkiev, A. I. Boldyrev and L. S. Wang, Angew. Chem., Int. Ed., 2007, 46, 4550 CrossRef CAS PubMed.
  21. B. B. Averkiev, D. Y. Zubarev, L. M. Wang, W. Huang, L. S. Wang and A. I. Boldyrev, J. Am. Chem. Soc., 2008, 130, 9248 CrossRef CAS PubMed.
  22. T. R. Galeev, W. L. Li, C. Romanescu, I. Černušák, L. S. Wang and A. I. Boldyrev, J. Chem. Phys., 2012, 137, 234306 CrossRef PubMed.
  23. B. B. Averkiev, L. M. Wang, W. Huang, L. S. Wang and A. I. Boldyrev, Phys. Chem. Chem. Phys., 2009, 11, 9840 RSC.
  24. L. Guennic, H. Jiao, S. Kahlal, J. Y. Saillard, J. F. Halet, S. Ghosh, M. Shang, A. M. Beatty, A. L. Rheingold and T. P. Fehlner, J. Am. Chem. Soc., 2004, 126, 3203 CrossRef PubMed.
  25. S. Ghosh, A. M. Beatty and T. P. Fehlner, J. Am. Chem. Soc., 2001, 123, 9188 CrossRef CAS PubMed.
  26. B. P. T. Fokwa and M. Hermus, Angew. Chem., Int. Ed., 2012, 51, 1702 CrossRef CAS PubMed.
  27. T. H. Pham and M. T. Nguyen, Phys. Chem. Chem. Phys., 2015, 17, 17335 RSC.
  28. T. B. Tai and M. T. Nguyen, J. Chem. Theory Comput., 2011, 7, 1119 CrossRef CAS PubMed.
  29. J. M. Tao, J. P. Perdew, V. N. Staroverov and G. E. Scuseria, Phys. Rev. Lett., 2003, 91, 146401 CrossRef PubMed.
  30. H. T. Pham, L. V. Duong, B. Q. Pham and M. T. Nguyen, Chem. Phys. Lett., 2013, 577, 32 CrossRef CAS.
  31. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  32. J. P. Perdew, K. Burke and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 16533 CrossRef CAS.
  33. I. Valencia and M. Castro, Phys. Chem. Chem. Phys., 2010, 12, 7545 RSC.
  34. H. Akeby and L. G. M. Pettersson, J. Mol. Spectrosc., 1993, 159, 17 CrossRef.
  35. M. Moskowits, D. P. DiLella and W. Limm, J. Chem. Phys., 1984, 80, 626 CrossRef.
  36. M. Doverstål, L. Karlsson, B. Lindgren and U. Sassenberg, Chem. Phys. Lett., 1997, 270, 273 CrossRef.
  37. M. Doverstål, B. Lindgren, U. Sassenberg, A. Arrington and M. D. Morse, J. Chem. Phys., 1992, 97, 7087 CrossRef.
  38. C. Cosse, M. Fouassier, T. Mejean, M. Tranwuille, D. P. DiLella and M. Moskowits, J. Chem. Phys., 1980, 73, 6076 CrossRef CAS.
  39. K. A. Gingerich, Faraday Symp. Chem. Soc., 1980, 14, 109 RSC.
  40. L. M. Russon, S. A. Heidecke, M. K. Birke, J. Conceicao, M. D. Morse and P. B. Armentrout, J. Chem. Phys., 1994, 100, 4747 CrossRef CAS.
  41. M. S. Gordon, J. S. Binkley, J. A. Pople, W. J. Pietro and W. J. Hehre, J. Am. Chem. Soc., 1983, 104, 2797 CrossRef.
  42. K. D. Dobbs and W. J. Hehre, J. Comput. Chem., 1987, 8, 861 CrossRef CAS.
  43. K. Raghavachari, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys., 1980, 72, 650 CrossRef.
  44. P. J. Hay, J. Chem. Phys., 1977, 66, 4377 CrossRef CAS.
  45. J. R. Cheeseman, J. A. Montgomery, T. Vreven, K. N. Kudin, J. C. Burant, J. Millam, et al., Gaussian 09 Revision: B.01, Gaussian, Inc., Wallingford, CT, 2009 Search PubMed.
  46. P. Lazzeretti, M. Malagoli and R. Zanasi, SYSMO package, Technical Report “Sistemi Informatici e Calcolo Parallelo”, CNR Italy (1991). Research Report number 1/67. Additional routines by P. W. Fowler, E. Steiner, R. W. A. Havenith and A. Soncini.
  47. R. W. A. Havenith and P. W. Fowler, Chem. Phys. Lett., 2007, 449, 347 CrossRef CAS.
  48. M. F. Guest, et al., Mol. Phys., 2005, 103, 719 CrossRef CAS , Gamess-UK.
  49. R. Zanasi, J. Chem. Phys., 1996, 105, 1460 CrossRef CAS.
  50. P. Lazzeretti, M. Malagoli and R. Zanasi, Chem. Phys. Lett., 1994, 220, 299 CrossRef CAS.
  51. K. B. Wiberg, Tetrahedron, 1968, 24, 1083 CrossRef CAS.
  52. I. Mayer, J. Comput. Chem., 2007, 28, 204 CrossRef CAS PubMed.
  53. R. L. John and B. Davis, Chem. Rev., 2002, 102, 2431 CrossRef.
  54. F. A. Cotton and D. G. Nocera, Acc. Chem. Res., 2000, 33, 483 CrossRef CAS PubMed.
  55. T. B. Tai, V. T. T. Huong and M. T. Nguyen, Top. Heterocycl. Chem., 2014, 38, 161 CAS.
  56. M. Kohout, F. R. Wanger and Y. Grin, Int. J. Quantum Chem., 2006, 106, 1499 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra04948f

This journal is © The Royal Society of Chemistry 2016