Next Article in Journal
Study of Anharmonicity in Zirconium Hydrides Using Inelastic Neutron Scattering and Ab-Initio Computer Modeling
Previous Article in Journal
Metal–Organic Frameworks as Versatile Platforms for Organometallic Chemistry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Metal-Rich Metallaboranes: Synthesis, Structures and Bonding of Bi- and Trimetallic Open-Faced Cobaltaboranes

1
Department of Chemistry, Indian Institute of Technology Madras, Chennai 600036, India
2
CNRS—Saint-Gobain—NIMS, IRL 3629, Laboratory for Innovative Key Materials and Structures (LINK), National Institute for Materials Science (NIMS), Tsukuba 305-0044, Japan
*
Authors to whom correspondence should be addressed.
Submission received: 22 March 2021 / Revised: 5 April 2021 / Accepted: 6 April 2021 / Published: 13 April 2021

Abstract

:
Synthesis, isolation, and structural characterization of unique metal rich diamagnetic cobaltaborane clusters are reported. They were obtained from reactions of monoborane as well as modified borohydride reagents with cobalt sources. For example, the reaction of [Cp*CoCl]2 with [LiBH4·THF] and subsequent photolysis with excess [BH3·THF] (THF = tetrahydrofuran) at room temperature afforded the 11-vertex tricobaltaborane nido-[(Cp*Co)3B8H10] (1, Cp* = η5-C5Me5). The reaction of Li[BH2S3] with the dicobaltaoctaborane(12) [(Cp*Co)2B6H10] yielded the 10-vertex nido-2,4-[(Cp*Co)2B8H12] cluster (2), extending the library of dicobaltadecaborane(14) analogues. Although cluster 1 adopts a classical 11-vertex-nido-geometry with one cobalt center and four boron atoms forming the open pentagonal face, it disobeys the Polyhedral Skeletal Electron Pair Theory (PSEPT). Compound 2 adopts a perfectly symmetrical 10-vertex-nido framework with a plane of symmetry bisecting the basal boron plane resulting in two {CoB3} units bridged at the base by two boron atoms and possesses the expected electron count. Both compounds were characterized in solution by multinuclear NMR and IR spectroscopies and by mass spectrometry. Single-crystal X-ray diffraction analyses confirmed the structures of the compounds. Additionally, density functional theory (DFT) calculations were performed in order to study and interpret the nature of bonding and electronic structures of these complexes.

Graphical Abstract

1. Introduction

Polyhedral cage expansion for the synthesis of large clusters has been the objective of boron-rich metallaborane and metallacarborane chemistry over the last six decades [1,2,3,4,5,6]. Molecular boron-rich clusters have found their place in many fields ranging from ceramics and polymers to boron neutron capture therapy and nanomaterials [7,8,9,10]. Those applications along with more fundamental studies of their specific and somewhat unique bonding and electronic structures have been the driving force towards the development of this chemistry [7,11,12]. In particular, following the pioneering work of Hawthorne et al. [13], several groups such as those of Evans [14], Welch [15], Xie [16,17], and ours [18,19,20] have contributed in enriching the library of the large 12–16 vertex single cage polymetalla(hetero)borane clusters over the years. The already extensive array of metallaborane clusters continues to grow in direct proportion with the advent of novel synthetic methods [9,11,17,21,22,23,24,25,26,27,28,29,30,31,32,33]. However, it is difficult to come up with an integrated scheme aimed at the synthesis of polyhedra bearing definite geometry and composition. This is largely due to the undefined and uncontrolled nature of cluster growth from metal synthons or small preformed clusters. Even the most sought-after strategies like condensation involving monoborane reagents, insertion or fragmentation involving borane or metal carbonyl fragments and cluster fusion [34] rarely exhibit metallaboranes with a high metal−to−boron ratio [35,36,37,38,39].
Among the several strategies aimed towards cluster expansion that have been conceived over the years, the reaction of cyclopentadienyl halides or hydrides with monoborane reagents such as BH3·THF (THF = tetrahydrofuran), [BHCl2·SMe2], or [LiBH4·THF], is a strategy that has been thoroughly optimized [40,41] and has proven to be very useful for yielding a variety of higher vertex metallaborane clusters adopting unique polyhedral cage geometries and structural features [41,42,43,44,45]. Although, in most cases, the metallaboranes are generated in small yields with a low metal−to−boron ratio, it was the strategy that involved pyrolyzing [Cp*RhCl2]2 in large excess of [BH3·THF], that allowed us to isolate 15- and 16-vertex rhodaborane clusters [18]. Although fused clusters [46] or single cage clusters bearing 5–11 vertices are reported with the cobalt system [20,46,47,48,49,50,51,52], isolation of analogous single cage supraicosahedral cobaltaboranes has not been successful up until now despite the fact that both rhodium and cobalt are group-9 metals. Indeed, our recent reports include the formation of a range of high-nuclearity cobaltaborane clusters, such as the hypoelectronic 9-vertex closo-[(Cp*Co)2B7H6OMe], hypoelectronic octadecahedral 11-vertex closo-[(Cp*Co)3B8H7R] (R = H or Me), and icosahedral 12-vertex closo-[(Cp*Co)3B8H8S] species obtained from the reaction of [Cp*CoCl]2 with [LiBH4·THF] and subsequent thermolysis with 2-mercaptobenzothiazole in toluene [50]. Our renewed quest for higher nuclearity cobaltaboranes made us look for modified synthetic strategies and although the objective of isolating a single cage supraicosahedral cobaltaborane was not achieved, the treatment of [Cp*CoCl]2 with [LiBH4·THF] followed by photolysis with excess BH3·THF afforded the 11-vertex-(hyper)nido-[(Cp*Co)3B8H10] cluster (1). Group-9 metallaboranes have continually monopolized the field not only for novelty in geometries but also for the possibility to draw on the open cage molecules like [(Cp*M)2B6H10] [M = Co and Rh] and [(Cp*Rh)2B8H12] for further reactivities [53]. We have ourselves contributed significantly to this area by enriching the library of mono- and dicobaltadecaborane(14) analogues in the past [50,51,52]. Most recently, we have explored and reported the chemistry of dicobaltaoctaborane(12) with chalcogenated borohydride, Li[BH2E3] [E = S, Se, or Te] which resulted in the characterization of novel chalcogenated analogues of dicobaltadecaborane(14) [54]. Careful analysis of the reaction between [(Cp*Co)2B6H10] and Li[BH2S3] led to the isolation of 2,4-[(Cp*Co)2B8H12] (2), a new dicobaltadecaborane(14) analogue. The main results are presented in this article.

2. Results and Discussion

2.1. Reaction of (Cp*CoCl)2 with [LiBH4·THF] and [BH3·THF]

The reaction of [Cp*CoCl]2 with an eight-fold excess of [LiBH4·THF] followed by photolysis with excess of BH3·THF at room temperature for 16 h led to the formation of a green crystalline solid in 10% yield which was characterized as a novel open-faced 11-vertex cobaltaborane, (hyper)nido-[(Cp*Co)3B8H10] (1). The reaction also yielded cobaltaboranes reported in the recent past like [(Cp*Co)2B6H10] and [(Cp*Co)B9H14] [52] (Scheme 1) and a few other compounds which could not be isolated due to lesser yields and transient natures. Preparative thin-layer chromatography (TLC) allowed separation of 1 from the reaction mixture and hence enabled the characterisation of the pure compound. Multinuclear NMR (see Supplementary Materials Figures S1−S4 for spectra) and IR spectroscopies and solid-state structure determination by X-ray diffraction analysis helped establish the composition and structure of 1. A detailed account of the spectroscopic and structural characterizations of 1 follows.
[(Cp*Co)3B8H10], 1. The room temperature 11B{1H} NMR of 1 shows the presence of four resonances between 71.4 and 19.7 ppm. The 1H{11B} NMR spectrum of 1 suggests the existence of two different Cp* environments appearing at 1.9 and 1.7 ppm with a relative ratio of 1:2. This is also corroborated by the 13C{1H} NMR spectrum. In addition to the chemical shifts corresponding to B-Ht protons, the 1H{11B} NMR spectrum also shows a sharp upfield resonance in the hydride region at δ = −20.43 ppm which may typically be arising due to two equivalent Co-H-B protons. The infrared (IR) spectrum shows broad bands around 2438 cm−1 owing to stretching frequencies of the terminal B-Ht bonds. Although the spectroscopic data indicated the formation of a large cluster, they were not sufficient to help conceptualize the actual identity of 1. A clear identification was obtained only after the crystal structure determination of 1 which was carried out with crystals obtained by slow evaporation of a concentrated hexane/CH2Cl2 (90:10) solution of 1 kept at −10 °C.
The molecular structure of 1 (Figure 1) adopts a classical 11-vertex nido core that can be formally generated by the removal of an apical vertex of the ubiquitous 12-vertex icosahedron (Figure 2). The molecule has two minor disordered components with atoms B8 and B8′, each with 50% occupancy. Despite the lack of precision in some of the bond distances and angles of the cluster owing to the restraints set to resolve this disorder, the crystallographic data reasonably allow a clear determination of the structure. Only one of the residues (with B8′) is discussed and shown in Figure 1. The five atoms of the open face (Co1, B4, B5, B7, and B8′) of the cluster are four-connected vertices while the remaining six ones are all five-connected vertices (Figure 1). This is in agreement with the NMR results discussed above which indicate that the higher chemical shifts in the 11B NMR spectrum corresponds to the high connected B atoms with the chemical shifts for B1 and B2 accidentally eclipsing at 71.4 ppm thus displaying four resonances in the ratio 2:2:2:2. A crystallographic mirror plane passing through the atoms Co1, B1, and B2 bisects the molecule into two symmetrical entities, thereby rendering the chemical environments of Cp* ligands attached to Co2 and Co3 as equivalent.
The polyhedral cage geometry of cluster 1 looks reminiscent of that of classical 11-vertex nido geometry of related clusters such as [C2B9H11]2−, [(Cp*Rh)3B8H9(OH)3] [55], [nido-7,8,9-PC2B8H11] [56], [8,8-η2-{η2-(BH3)·dppm}-nido-8,7-RhSB9H10] [57], [9,9-η2-{η2-(BH3)·dppm}-nido-9,7,8-RhC2B8H11] [57], and [(μ-PPh2)(PPh3)2Pt2B9H6(OiPr)3] [58] for instance which all possess 13 skeletal electron pairs (SEP), i.e., (n + 2) SEPs if n is the number of occupied vertices according to the Polyhedral Skeletal Electron Pair Theory (PSEPT) [59]. Interestingly, however, with only 12 skeletal electron pairs (SEPs), cluster 1 is two electrons short of the formal (n + 2) SEP count expected for nido clusters as required by the PSEPT [59]. As stated earlier, in addition to the BHt and Cp* protons, the 1H{11B} NMR spectrum shows the presence of only a single resonance for two bridging Co-H-B protons. The presence of two B-H-B protons which would have afforded a nido species with the “expected” electron count could not be traced from 1H NMR.
To the best of our knowledge cluster 1 is the first example of electron-poor 11-vertex−12-SEP (hyper)nido species of this kind in the metallaborane cluster domain. Nevertheless, isoelectronic and isostructural examples are known in metallacarborane chemistry. Indeed, cluster 1 can be compared both electronically and structurally to some 11-vertex−12-SEP nido clusters such as [9,9-L2-nido-9,7,8-MC2B8H11] (I: L = PPh3, M = Ir; II: L = PEt3, M = Rh) [60] and [8,8-(PPh3)2-nido-8,7-RhSB9H10] (III) [61,62] (Chart 1). 11-vertex−12-SEP metalla(hetero)borane compounds exist such as [1-PPh3-{1,3-(µ-dppm)}-isonido-1,2-RhSB9H8] [IV; dppm = CH2(PPh2)2] [63,64], or [1,1,1-(H)(PMe3)2-isonido-1,2-IrSB9H9] (V) [65] (Figure 3), but they are not isostructural with 1. While the latter is generated from the icosahedron by removing one vertex (Figure 2), the geometry reported for the 11-vertex isonido clusters [63,64,65] is derived from an 11-vertex octadecahedron by the removal of one connectivity of the apical vertex thereby generating a distorted quadrilateral open face (D; top of Figure 3).
Density functional theory (DFT) calculations were carried out at the B3LYP/LANL2DZ level of theory (see computational details in Section 3) on model compound 1′ (Cp analogue of 1) to analyse the electronic structure and nature of bonding in the electron deficient molecule 1. Optimized bond distances in 1′ are in accordance with those experimentally measured for the solid-state structure of 1 (Table S1). The molecular orbital (MO) analysis of 1′ (Table S2) shows a large energy gap of 2.79 eV between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) in agreement with the electron count of 12 SEPs and the fact that only two bridging hydrogen atoms are present in the structure. The HOMO of 1′ is mainly localized on the metal centres and the five-connected inner-core boron atoms, whereas the LUMO is predominantly localized on the metal centres with considerable participation of the boron atoms of the open face of the cluster (Figure 4). We may wonder if cluster 1 could be reduced and reach the expected count of 13 SEPs to obey the PSEPT. It turns out that the LUMO + 1 is energetically close to the LUMO, separated only by 0.4 eV. This indicates that reduction of 1 is unlikely.

2.2. Reactivity of [(Cp*Co)2B6H12] with Li[BH2S3]

Reactions of the open-cage dicobaltaoctaborane(12) analogue [(Cp*Co)2B6H12] with chalcogenated borohydride Li[BH2E3] (E = S, Se, or Te) which were recently carried out in our group, resulted in the formation of novel chalcogenated analogues of dicobaltadecaborane(14), namely [(Cp*Co)2B6E2H6] (E = S, Se, or Te) [54]. The room temperature reaction between [(Cp*Co)2B6H12] and Li[BH2S3] yielded another compound which could not be isolated earlier due to less yield. Performing the reaction at a larger scale followed by careful analysis of the reaction mixture led to the isolation of 2,4-[(Cp*Co)2B8H12] (2), i.e., a novel dicobaltadecaborane(14) analogue (Scheme 2). This compound was characterized by 1H, 11B{1H} and 13C{1H} NMR (Figures S5−S8) and IR spectroscopies, mass spectrometry (Figure S9), and single-crystal X-ray diffraction analysis.
Dicobaltadecaborane(14) analogue, 2. Cluster 2 was isolated as a yellow solid in 15% yield. The mass spectrometric data for 2 showed molecular ion peaks at m/z 487.2692 which corresponds to the molecular formula [C20H42B8Co2]+. The 1B{1H} NMR showed chemical shifts at δ = 26.6, 11.4, and 5.6 ppm, whereas 1H NMR showed a single chemical shift for 30 Cp* protons at δ = 1.69 ppm, and additionally, chemical shifts at δ = 4.38, 2.94, and 2.31 ppm which correspond to B-Ht protons. A distinct up-field resonance was observed at δ = −4.25 ppm which may be assigned to four equivalent B-H-B bridging protons. The chemical shifts displayed in the 13C{1H} NMR spectrum also corroborated with a single Cp* environment in the molecule. The infrared (IR) spectrum showed broad bands centered around 2493 cm−1 owing to stretching frequencies of the terminal B-H bonds.
Slow evaporation of a concentrated hexane/CH2Cl2 (80:20) solution of 2 kept at −10 °C yielded crystals suitable for X-ray diffraction analysis. The solid-state X-ray structure of 2 (Figure 5), determined by single-crystal X-ray diffraction measurements clearly shows that it is a 10-vertex open nido-dicobaltaborane which is in absolute agreement with the spectroscopic as well as mass spectrometric data. It is apparent from its solid state X-ray structure that compound 2 adopts the same structural arrangement as the symmetric dimetalla nido-2,4-[(Cp*M)2B8H12] (M = Rh and Ir) [27,55] previously reported, analogues of decaborane(14). In the structure, a crystallographic mirror plane including the B6-B8 vector bisects the molecule into two equivalent {CoB3} units. As a result, the two Co atoms that reside on either side of the crystallographic mirror plane are chemically equivalent and hence correspond to the presence of single Cp* environment in the 1H and 13C{1H} NMR spectra. The bridging hydrogen atoms lie on the open face of the six-membered ring and their presence is confirmed both by 1H NMR and X-ray diffraction studies. We noted that the average B-B bond length in 2 is somewhat shorter as compared to the Rh and Ir analogues [27,55] and follows the trend (Co < Rh < Ir) down the triad. A {Cp*Co} unit being isolobal with a BH unit, the cluster is isostructural and isoelectronic with nido-decaborane(14) which can be generated from the removal of the six-connected apical vertex of an octadecahedron (see bottom of Figure 3). Indeed, 2 can be regarded as a 10-vertex nido metallaborane cluster with 12 SEPs, as expected according to the PSEPT. 2 is then a new member of the series of ten-vertex−12-SEP polyhedral borane, carborane, or metallaborane clusters having the same SEP count and geometry. We have tried to compare its structural data and chemical shift values with a set of comparable cobaltaboranes with the formal electron count of 12 SEPs (Table 1). The unique feature of this reaction here (Scheme 2) is the insertion of two BH units in octaborane(12) that leads to cluster growth from 8 to 10 vertices.
In order to acquire an insight into the structure and bonding of 2, we further did DFT calculations at the B3LYP/LANL2DZ level of theory (see computational details in Section 3 on model 2′ (Cp analogue of 2) as well as on its group-9 metallaborane model analogues 3′ (Cp analogue of 2,4-[(Cp*Rh)2B8H12] and 4′ (Cp analogue of 2,4-[(Cp*Ir)2B8H12]) for comparison. The optimized geometrical parameters obtained for 2′ compare fairly well with the corresponding crystallographically characterized data of 2 (Table S1). Additionally, the gauge-including atomic orbitals-density functional theory (GIAO-DFT calculated 11B chemical shift values agree well with those observed experimentally (Table 2) and are of sufficient accuracy to confirm assignments that are consistent with the empirical ones based simply on boron connectivity and environment (see above).
A molecular orbital analysis suggests a significant HOMO-LUMO energy gap (>4 eV) for the three compounds in the order 2′ < 3′ < 4′ (Figure 6). Inspection of the nodal properties of the frontier MOs reveals that HOMOs weigh heavily on the respective metal centres, while the LUMOs are delocalized both on the metal centres and the five-connected inner-core boron atoms. Metal−metal Wiberg bond indices computed for 2′, 3′, and 4′ are very weak and close to each other (ca. 0.06), reflecting the nonappearance of bonding interaction between the metal centres in such species.

3. Materials and Methods

3.1. General Methods and Instrumentation

All of the experiments were conducted under an argon atmosphere using standard Schlenk and glove-box techniques. Solvents were distilled prior to use under argon. Cp*H, CoCl2, n-BuLi in hexane, LiBH4 (2.0 M in THF), BH3∙THF, and S powder were used as received (Sigma Aldrich, Bangalore, India). (Cp*CoCl)2 [67,68], [(Cp*Co)2B6H10] [52], and Li[BH2S3] [69,70] were synthesized as per literature. Thin-layer chromatography, using silica gel TLC plates by Merck supported with 250-μm diameter aluminum, was done to separate the reaction mixtures. A 500 MHz Bruker FT-NMR spectrometer (Bruker, Billerica, MA, USA), JASCO FT/IR-4100 (JASCO, India) and Agilent Technologies: 6545-Q-TOF LC/MS (Agilent, Santa Clara, CA, USA) spectrometer were used respectively to record NMR, Infrared (IR) spectra, and ESI (Electronspray Ionization) mass. In 1H NMR spectra, the chemical shifts for residual protons in the deuterated solvents were used as reference (CDCl3, δ = 7.26 ppm, C6D6, δ = 7.16) while a sealed tube containing [Bu4N(B3H8)] [71] in benzene-d6 (δB, ppm, −30.07) was used as an external reference for 11B{1H} NMR spectra. The light source for the photoreaction described in this report was Oriel Instrument with a 500 W Hg/Xe arc lamp equipped with liquid and colored glass filters, irradiating at 280–780 nm.
Synthesis of 1: [Cp*CoCl]2 (0.1 g, 0.22 mmol) was suspended in toluene (15 mL) in a flame-dried Schlenk tube and was cooled to −78 °C, [LiBH4∙THF] (0.6 mL, 1.20 mmol) was added to it and the reaction mixture was allowed to warm slowly to room temperature and was left to stir for another hour. [BH3∙THF] (3 mL) was added to the reaction mixture and the resultant mixture was slowly warmed to room temperature was allowed to stir for 30 min, and finally kept for photolysis at room temperature for 12 h. After completion of the reaction (monitored by TLC), the solvent was removed under vacuum and the residue was dissolved in hexane and passed through Celite. The solvent was evaporated under vacuum and the residue was purified using preparative silica-gel TLC plates by eluting with hexane/CH2Cl2 (80:20) mixture that yielded green solid 1 (0.010 g, 10%) and known red solids [(Cp*Co)2B6H10] (0.025 g, 24%), [(Cp*Co)3B4H4] (0.018 g, 19%), and a few compounds which could not be isolated. The known cobaltaboranes were characterized with reference to their spectroscopic data reported earlier.
1: 11B NMR (160 MHz, C6D6, 22 °C): δ = 71.5, 31.4, 24.5, 19.7 ppm; 1H{11B} NMR (500 MHz, C6D6, 22 °C): δ = 8.08 (br, 4H, BHt), 5.16 (s, 2H, BHt), 4.08 (s, 2H, BHt), 1.94 (s, 15H, Cp*), 1.70 (s, 30H, Cp*) −20.43 (br, 2H, Co-H-B) ppm; 13C NMR (100 MHz, C6D6, 22 °C) = 96.0, 93.4 (C5Me5), 10.4, 9.6 ppm (C5Me5); IR (dichloromethane): = 2438 (BHt) cm−1.
Synthesis of 2: In a flame-dried Schlenk tube, [(Cp*Co)2B6H10] (0.05 g, 0.107 mmol) was suspended in 5 mL of dry THF at room temperature. A freshly prepared solution of Li[BH2S3] in THF (10 mL, 0.428 mmol) was added dropwise to the red solution of [(Cp*Co)2B6H10]. The reaction mixture was allowed to stir at room temperature for 24 h. The solvent was dried under vacuum and the residue was extracted into n-hexane/CH2Cl2 (85:15 v/v) mixture and passed through Celite. After removal of the solvent, the residue was subjected to chromatographic work-up by using TLC plates. Elution with an n-hexane/CH2Cl2 (85:15 v/v) mixture yielded the air-stable yellow solid (Compound 2) (0.020 g, 38%) and the known compound [(Cp*Co)2B6H6S2] (0.016 g, 28 %) [57].
2: MS (ESI+): m/z calculated for C20H42B8Co2+ 487.2747, found 487.2692; 11B NMR (160 MHz, CDCl3, 22 °C): δ = 26.6, 11.4, 5.6 ppm; 1H NMR (500 MHz, CDCl3, 22 °C): δ = 4.38 (s, 2H, BHt), 2.94 (s, 2H, BHt), 2.31 (s, 4H, BHt), 1.69 (s, 30H, Cp*), −4.25 (br, 4H, B-H-B) ppm; 13C NMR (100 MHz, CDCl3, 22 °C) = 97.0 (C5Me5), 9.8 ppm (C5Me5); IR (dichloromethane): = 2493 (BHt) cm−1.

3.2. Single Crystal X-ray Diffraction Analysis

Crystallographic and structural refinement data for 1 and 2 are given in Table 3. Bruker AXS Kappa APEXII CCD diffractometer, (Bruker, Billerica, MA, USA) with graphite monochromatic Mo Kα (λ = 0.71073 Å) radiation at 296 K was used to collect and integrate crystal diffraction data. Heavy atom methods like SHELXT-2014 and SHELXS-97 [72,73,74] were used to solve the structures while SHELXL-2018 and SHELXL-2014 [73] were used for structure-refinements, respectively. Olex2 [75] was used for drawing molecular structures.

3.3. Computational Details

Gaussian 09 program package [76] was used for geometry optimizations of all model clusters. Computations were done with the Cp ligands in place of Cp* to reduce the computational effort (for optimized model clusters see Figures S10–S13). Gas-phase optimizations without any solvent effect or symmetry constraints, were done with the hybrid Becke–Lee–Yang–Parr (B3LYP) functional [77] in combination with the LANL2DZ (Los Alamos National Laboratory 2 Double-Zeta) basis set [78] for all the atoms from the EMSL (Environmental Molecular Sciences Laboratory) Basis Set Exchange Library [79]. Frequency calculations were carried out for all structures to check the nature of the stationary states and the absence of any imaginary frequency to confirm that the optimized geometries were genuine minima on the potential energy hypersurface. NMR chemical shifts were computed by employing the gauge-including atomic orbitals (GIAOs) method [80,81,82] using the optimized geometries at the B3LYP/LANL2DZ level. Chemical shifts corresponding to 11B NMR were computed in relation to B2H6 (B3LYP shielding constant for 11B NMR: 83.6 ppm) and were then converted to the standard [BF3·OEt2] scale by adding 16.6 ppm (the experimental δ (11B) value of B2H6) to the computed values. 1H NMR chemical shift calculations were done with TMS (SiMe4) as the internal standard. Chemcraft [83] was used for generating orbital graphics and optimized structure plots. Natural Bond Orbital (NBO) analyses [84] allowed for arrival of the Wiberg Bond Indices (WBI) [85].

4. Conclusions

Results described in this article have demonstrated the scope of new methodologies for the synthesis of high-nuclearity cobaltaborane clusters. Firstly, one of the new routes described here has led to the isolation of a novel and unprecedented 11-vertex–12-SEP disobedient cobaltaborane cluster with a (hyper)nido M3B8 core. The second route has allowed for synthesis and characterization of a classic example of sequentially generated higher nuclearity cluster with a 10-vertex–12-SEP nido core via cluster growth on a preformed template. Indeed, cobaltaboranes and cobaltaheteroboranes constitute a showcase for dynamic and unique structural features and as a result, we are always on the lookout for possibilities of generating novel sophisticated cobaltaborane frameworks. Amongst the several synthetic strategies that are recognized in this domain, the applicability of the ones reported here may be vital and might lead to the generation of sophisticated high-nuclearity metallaborane frameworks of other group-9 metals. The unusual formation of 1 and the open-faced geometries of both 1 and 2 should imply enhanced reactivity to such species. Further efforts are currently devoted to that purpose.

Supplementary Materials

Some supplementary information is available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/inorganics9040028/s1. It contains 1H, 11B{1H}, 13C{1H} NMR and mass spectra (Figures S1–S9); CIF and checkCIF files as well as the xyz coordinates of the DFT optimized model clusters (Figures S10—S13), selected bond parameters for clusters 1′ and 2′ and their Wiberg bond indices (WBI) (Table S1) as well as calculated HOMO and LUMO energy levels and HOMO-LUMO gaps of 14′ (Table S2).

Author Contributions

K.P. and C.N. formulated and planned the experiments; K.P. and C.N. carried out the synthesis and spectroscopic analyses and discussed the results with S.G. and J.-F.H.; K.P. and C.N. organized the manuscript with guidance from S.G. and J.-F.H.; S.G. supervision, S.G. project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by CSIR (Scheme No. 01 (2939)/18/EMR-II), New Delhi, India.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Accession codes: 2043679 (1), 2060786 (2). Crystallographic data can be acquired for free from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44-1223-336033.

Acknowledgments

K.P. thanks IIT Madras and C.N. thanks the Department of Science and Technology, India, DST-INSPIRE for research fellowships. The Department of Chemistry, IIT Madras is thankfully acknowledged for X-ray support. The authors thank IIT Madras for computational facilities and Babu Varghese and Venkatachalam Ramkumar for Single Crystal X-ray Diffraction data analyses.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lipscomb, W.L. Boron Hydrides; Benjamin: New York, NY, USA, 1963. [Google Scholar]
  2. Muetterties, E.L. Boron Hydride Chemistry; Academic Press: New York, NY, USA, 1975. [Google Scholar]
  3. Grimes, R.N. Carboranes, 3rd ed.; Elsevier: Oxford, UK, 2016. [Google Scholar]
  4. King, R.B. Three-Dimensional Aromaticity in Polyhedral Boranes and Related Molecules. Chem. Rev. 2001, 101, 1119–1152. [Google Scholar] [CrossRef]
  5. Jemmis, E.D. Overlap control and stability of polyhedral molecules. closo-Carboranes. J. Am. Chem. Soc. 1982, 104, 7017–7020. [Google Scholar] [CrossRef]
  6. Hosmane, N.S.; Maguire, J.A. Metallacarboranes of d- and f-Block Metals. In Comprehensive Organometallic Chemistry III; Mingos, D.M.P., Crabtree, R.H., Eds.; Pergamon: New York, NY, USA, 2007; Chapter 3.05; Volume 3, pp. 175–264. [Google Scholar]
  7. Hosmane, N.S. Boron Science: New Technologies and Applications; CRC Press, Taylor and Francis Group: Boca Raton, FL, USA, 2011. [Google Scholar]
  8. Saxena, A.K.; Hosmane, N.S. Recent advances in the chemistry of carborane metal complexes incorporating d- and f-block elements. Chem. Rev. 1993, 93, 1081–1124. [Google Scholar] [CrossRef]
  9. Grimes, R.N. Transition Metal Metallacarbaboranes. In Comprehensive Organometallic Chemistry II; Abel, E.W., Stone, F.G.A., Wilkinson, G., Eds.; Pergamon Press: Oxford, UK, 1995; Volume 1, pp. 373–430. [Google Scholar]
  10. Housecroft, C.E. Boron Atoms in Transition Metal Clusters. Adv. Organomet. Chem. 1991, 33, 1–50. [Google Scholar]
  11. Fehlner, T.P.; Halet, J.-F.; Saillard, J.-Y. Molecular Clusters. A Bridge to Solid-State Chemistry; University Press: Cambridge, UK, 2007. [Google Scholar]
  12. Zhang, J.; Xie, Z. Synthesis, structure, and reactivity of 13- and 14-Vertex Carboranes. Acc. Chem. Res. 2014, 47, 1623–1633. [Google Scholar] [CrossRef]
  13. Hawthorne, M.F.; Dunks, G.B.; McKown, M.M. Probable formation of 13-atom polyhedral complexes containing B10C2H122− and cobalt. J. Am. Chem. Soc. 1971, 93, 2541–2543. [Google Scholar] [CrossRef]
  14. Evans, W.J.; Hawthorne, M.F. Synthesis of fourteen vertex metallocarboranes by polyhedral expansion. J. Chem. Soc. Chem. Commun. 1974, 38–39. [Google Scholar] [CrossRef]
  15. Burke, A.; Ellis, D.; Giles, B.T.; Hodson, B.E.; Macgregor, S.A.; Rosair, G.M.; Welch, A.J. Beyond the Icosahedron: The First 13-Vertex Carborane. Angew. Chem. Int. Ed. 2003, 42, 225–228. [Google Scholar] [CrossRef]
  16. Deng, L.; Chan, H.-S.; Xie, Z. Synthesis, Reactivity, and Structural Characterization of a 14-Vertex Carborane. Angew. Chem. Int. Ed. 2005, 44, 2128–2131. [Google Scholar] [CrossRef] [PubMed]
  17. Deng, L.; Zhang, J.; Chan, H.-S.; Xie, Z. Synthesis and Structure of 14- and 15-Vertex Ruthenacarboranes. Angew. Chem. Int. Ed. 2006, 45, 4309–4313. [Google Scholar] [CrossRef] [PubMed]
  18. Roy, D.K.; Bose, S.K.; Anju, R.S.; Mondal, B.; Ramkumar, V.; Ghosh, S. Boron beyond the Icosahedral Barrier: A 16-Vertex Metallaborane. Angew. Chem. Int. Ed. 2013, 52, 3222–3226. [Google Scholar] [CrossRef]
  19. Roy, D.K.; Mondal, B.; Shankhari, P.; Anju, R.S.; Geetharani, K.; Mobin, S.M.; Ghosh, S. Supraicosahedral Polyhedra in Metallaboranes: Synthesis and Structural Characterization of 12-, 15- and 16-Vertex Rhodaboranes. Inorg. Chem. 2013, 52, 6705–6712. [Google Scholar] [CrossRef]
  20. Roy, D.K.; Ghosh, S.; Halet, J.-F. Beyond the Icosahedron: The Quest for High-nuclearity Supraicosahedral Metallaboranes. J. Clust. Sci. 2014, 25, 225–237. [Google Scholar] [CrossRef]
  21. Onak, T. Polyhedral Carboranes. In Comprehensive Organometallic Chemistry II; Abel, E.W., Stone, F.G.A., Wilkinson, G., Eds.; Pergamon Press: Oxford, UK, 1995; Chapter 6; Volume 1, pp. 217–255. [Google Scholar]
  22. Hosmane, N.S.; Siebert, W. (Eds.) Advances in Boron Chemistry; Royal Society of Chemistry: Cambridge, UK, 1997; p. 349. [Google Scholar]
  23. Ghosh, S.; Rheingold, A.L.; Fehlner, T.P. Metallaboranes of the Earlier Transition Metals. An Arachno Nine-vertex, Nine-Skeletal Electron Pair Rhenaborane of Novel Shape: Importance of Total Vertex Connectivities in Such Systems. Chem. Commun. 2001, 10, 895–896. [Google Scholar] [CrossRef]
  24. Ghosh, S.; Lei, X.; Shang, M.; Fehlner, T.P. Role of the Transition Metal in Metallaborane Chemistry. Reactivity of [(Cp*ReH2)2B4H4] with BH3·THF, CO, and [Co2(CO)8]. Inorg. Chem. 2000, 39, 5373–5382. [Google Scholar] [CrossRef] [PubMed]
  25. Ghosh, S.; Noll, B.C.; Fehlner, T.P. Borane Mimics of Classic Organometallic Compounds: [(Cp*Ru)B8H14(RuCp*)]0,+, Isoelectronic Analogues of Dinuclear Pentalene Complexes. Angew. Chem. Int. Ed. 2005, 44, 6568–6571. [Google Scholar] [CrossRef]
  26. Ghosh, S.; Fehlner, T.P.; Noll, B.C. Condensed metallaborane clusters: Synthesis and structure of Fe2(CO)65-C5Me5RuCO)(η5-C5Me5Ru)B6H10. Chem. Commun. 2005, 3080–3082. [Google Scholar] [CrossRef] [PubMed]
  27. Ghosh, S.; Noll, B.C.; Fehlner, T.P. Expansion of Iridaborane Clusters by Addition of Monoborane. Novel Metallaboranes and Mechanistic Detail. Dalton Trans. 2008, 3, 371–378. [Google Scholar] [CrossRef] [PubMed]
  28. Geetharani, K.; Bose, S.K.; Pramanik, G.; Saha, T.K.; Ramkumar, V.; Ghosh, S. An Efficient Route to Group 6 and 8 Metallaborane Compounds: Synthesis of arachno-[Cp*Fe(CO)B3H8] and closo-[(Cp*M)2B5H9] (M = Mo, W). Eur. J. Inorg. Chem. 2009, 2009, 1483–1487. [Google Scholar] [CrossRef]
  29. Dhayal, R.S.; Sahoo, S.; Reddy, K.H.K.; Mobin, S.M.; Jemmis, E.D.; Ghosh, S. Vertex-Fused Metallaborane Clusters: Synthesis, Characterization and Electronic Structure of [(η5-C5Me5Mo)3MoB9H18]. Inorg. Chem. 2010, 49, 900–904. [Google Scholar] [CrossRef]
  30. Geetharani, K.; Bose, S.K.; Sahoo, S.; Varghese, B.; Mobin, S.M.; Ghosh, S. Cluster Expansion Reactions of Group 6 and 8 Metallaboranes Using Transition Metal Carbonyl Compounds of Groups 7-9. Inorg. Chem. 2011, 50, 5824–5832. [Google Scholar] [CrossRef] [PubMed]
  31. Chakrahari, K.K.V.; Thakur, A.; Mondal, B.; Dhayal, R.S.; Ramkumar, V.; Ghosh, S. A Close-packed Boron-rich 11-vertex Molybdaborane with Novel Geometry. J. Organomet. Chem. 2012, 710, 75–79. [Google Scholar] [CrossRef]
  32. Roy, D.K.; Barik, S.K.; Mondal, B.; Varghese, B.; Ghosh, S. A Novel Heterometallic μ9-Boride Cluster: Synthesis and Structural Characterization of [(η5-C5Me5Rh)2{Co6(CO)12}(μ-H)(BH)B]. Inorg. Chem. 2014, 53, 667–669. [Google Scholar] [CrossRef] [PubMed]
  33. De, A.; Zhang, Q.-F.; Mondal, B.; Cheung, L.F.; Kar, S.; Saha, K.; Varghese, B.; Wang, L.-S.; Ghosh, S. [(Cp2M)2B9H11] (M = Zr or Hf): Early Transition Metal ‘guarded’ Heptaborane with Strong Covalent and Electrostatic Bonding. Chem. Sci. 2018, 9, 1976–1981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Bould, J.; Rath, N.P.; Barton, L. Metallaborane heteroatom incorporation reactions: Metallacarboranes, metallathiaboranes, and an iridaazaborane from iridanonaborane precursors. Organometallics 1996, 15, 4916–4929. [Google Scholar] [CrossRef]
  35. Ferguson, G.; Lough, A.J.; Faridoon; McGrath, M.N.; Spalding, T.R.; Kennedy, J.D.; Fontaine, X.L.R. Metallaheteroborane chemistry. Part 7. Synthesis, crystal structure, and characterisation of two dinuclear rhodatelluraboranes, [{(PPh3)2RhTeB10H10}2] and [(PPh3)(CO)Rh2Te2B20H20]. J. Chem. Soc. Dalton Trans. 1990, 6, 1831–1839. [Google Scholar] [CrossRef]
  36. Ferguson, G.; Gallagher, J.F.; McGrath, M.; Sheehan, J.P.; Spalding, T.R.; Kennedy, J.D. Metallaheteroborane chemistry. Part 11. Selective syntheses of the palladium heteroborane complexes [2,2-(PR3)2-closo-2,1-PdEB10H10] (R3 = Ph3, MePh2 or Me2Ph; E = Se or Te) and [2-X-2-(PPh3)-closo-2,1-PdTeB10H9(PPh3)] (X = Cl, Br, I, CN, SCN or O2CMe). J. Chem. Soc. Dalton Trans. 1993, 1, 27–34. [Google Scholar] [CrossRef]
  37. Norman, N.C.; Orpen, A.G.; Quayle, M.J.; Rice, C.R. Diborane(4) compounds incorporating thio- and seleno-carboranyl groups. New J. Chem. 2000, 24, 837–839. [Google Scholar] [CrossRef]
  38. Hammerschmidt, A.; Doch, M.; Putz, S.; Krebs, B. Na2[B18Se16]: The first 3D polymeric selenoborato-closo-dodecaborate. Z. Anorg. Allg. Chem. 2005, 631, 1125–1128. [Google Scholar] [CrossRef]
  39. Ferguson, G.; Gallagher, J.F.; Kennedy, J.D.; Kelleher, A.M.; Spalding, T.R. Pentahapto-bonded gold heteroborane clusters [3-(R3P)-closo-2,1-AuTeB10H10] and [3-(R3P)-closo-3,1,2-AuAs2B9H9]. Dalton Trans. 2006, 7, 2133–2139. [Google Scholar] [CrossRef] [PubMed]
  40. Bose, S.K.; Geetharani, K.; Ramkumar, V.; Varghese, B.; Ghosh, S. Chemistry of vanadaboranes: Synthesis, structures and characterization of organ vanadium sulfide clusters with disulfide linkage. Inorg. Chem. 2010, 49, 2881–2888. [Google Scholar] [CrossRef]
  41. Ghosh, S.; Noll, B.C.; Fehlner, T.P. Synthesis and characterization of [exo-BH2(Cp*M)2B9H14] (M = Ru, Re), and the conversion of the ruthenaborane into [(Cp*Ru)2B10H16] with an open cluster framework based on a capped truncated tetrahedron. Angew. Chem. Int. Ed. 2005, 44, 2916–2918. [Google Scholar] [CrossRef] [PubMed]
  42. Ghosh, S.; Shang, M.; Fehlner, T.P. Comparison of the geometric and molecular orbital structures of (Cp*Cr)2B4H8 and (Cp*Re)2B4H8, Cp* = η5-C5Me5. Structural consequences of delocalized electronic unsaturation in a metallaborane cluster. J. Organomet. Chem. 2000, 614, 92–98. [Google Scholar] [CrossRef]
  43. Geetharani, K.; Bose, S.K.; Varghese, B.; Ghosh, S. From metallaborane to borylene complexes: Syntheses and structures of triply bridged ruthenium and tantalum borylene complexes. Chem. Eur. J. 2010, 16, 11357–11366. [Google Scholar] [CrossRef] [PubMed]
  44. Bose, S.K.; Ghosh, S.; Noll, B.C.; Halet, J.-F.; Saillard, J.-Y.; Vega, A. Linked and fused tungstaborane clusters: Synthesis, characterization and electronic structures of bis-{(η5-C5Me5W)2B5H8}2 and (η5-C5Me5W)2{Fe(CO)3}nB6-nH10-n, n = 0,1. Organometallics 2007, 26, 5377–5385. [Google Scholar] [CrossRef]
  45. Ghosh, S.; Lei, X.; Cahill, C.L.; Fehlner, T.P. Symmetrical scission of the coordinated tetraborane in (Cp*ReH2)B4H4 on CO addition and reassociation of the coordinated diboranes on H2 loss. Angew. Chem. Int. Ed. 2000, 39, 2900–2902. [Google Scholar] [CrossRef]
  46. Zafar, M.; Kar, S.; Nandi, C.; Ramalakshmi, R.; Ghosh, S. Cluster Fusion: Face-Fused Macropolyhedral Tetracobaltaboranes. Inorg. Chem. 2019, 58, 47–51. [Google Scholar] [CrossRef]
  47. Venable, T.L.; Sinn, E.; Grimes, R.N. Cobaltaborane analogs of decaborane (B10H14). Crystal and molecular structures of 6-[η5-C5(CH3)5]CoB9H13, 6,9-[η5-C5(CH3)5]2]2Co2B8H12, 5,7-[η5-C5(CH3)5]5]2Co2 B8H12, and 6-Cl-5,7-[η5-C5(CH3)5]2Co2B8H11]. Inorg. Chem. 1982, 21, 895–904. [Google Scholar] [CrossRef]
  48. Venable, T.L.; Grimes, R.N. (Pentamethylcyclopentadienyl)- cobaltaboranes derived from the octah dropentaborate (B5H8-) and tetradecahydrononabaorate (B9H14) ions: Studies in synthesis and structure. Inorg. Chem. 1982, 21, 887–892. [Google Scholar] [CrossRef]
  49. Pipal, J.R.; Grimes, R.N. Crystal structure of a tetracobalt tetraboron cluster, (η5-C5H5)4Co4B4H4. Structural patterns in eight-vertex polyhedra. Inorg. Chem. 1979, 18, 257–263. [Google Scholar] [CrossRef]
  50. Chakrahari, K.K.; Sharmila, D.; Barik, S.K.; Mondal, B.; Varghese, B.; Ghosh, S. Hypoelectronic metallaboranes: Synthesis, structural characterization and electronic structures of metal-rich cobaltaboranes. J. Organomet. Chem. 2014, 749, 188–196. [Google Scholar] [CrossRef]
  51. Barik, S.K.; Roy, D.K.; Sharmila, D.; Ramalakshmi, R.; Chakrahari, K.K.; Mobin, S.M.; Ghosh, S. Synthesis, Characterization and Electronic Structures of Rh and Co analogs of Decaborane-14. Proc. Natl. Acad. Sci. India Sect. A 2014, 84, 121–130. [Google Scholar] [CrossRef]
  52. Barik, S.K.; Roy, D.K.; Ghosh, S. Chemistry of group 9 dimetallaborane analogues of octaborane(12). Dalton Trans. 2015, 44, 669–676. [Google Scholar] [CrossRef] [Green Version]
  53. Roy, D.K.; Anju, R.S.; Varghese, B.; Ghosh, S. Reactivity of dirhodium analogs of octaborane-12 and decaborane-14 towardstransition-metal moieties. Organometallics 2013, 32, 1964–1970. [Google Scholar] [CrossRef]
  54. Nandi, C.; Kar, S.; Zafar, M.; Kar, K.; Ghosh, S. Chemistry of Dimetallaoctaborane(12) with Chalcogen-Based Borate Ligands: Obedient versus Disobedient Clusters. Inorg. Chem. 2020, 59, 3537–3541. [Google Scholar] [CrossRef]
  55. Roy, D.K.; Bose, S.K.; Anju, R.S.; Ramkumar, V.; Ghosh, S. Synthesis and Structure of Dirhodium Analogue of Octaborane-12 and Decaborane-14. Inorg. Chem. 2012, 51, 10715–10722. [Google Scholar] [CrossRef]
  56. Stibr, B.; Holub, J.; Bakardjiev, M.; Hnyk, D.; Tok, O.L.; Milius, W.; Wrackmeyer, B. Phosphacarborane Chemistry: The Synthesis of the Parent Phosphadicarbaboranes nido-7,8,9-PC2B8H11 and [nido-7,8,9-PC2B8H10], and Their 10-Cl Derivatives-Analogs of the Cyclopentadienide Anion. Eur. J. Inorg. Chem. 2002, 2002, 2320–2326. [Google Scholar] [CrossRef]
  57. Volkov, O.; Macias, R.; Rath, N.P.; Barton, L. Phosphine−Boranes as Bidentate Ligands: Formation of [8,8-η2-{η2-(BH3)-dppm}-nido-8,7-RhSB9H10] and [9,9-η2-{η2-(BH3)-dppm}-nido-9,7,8-RhC2B8H11] from [8,8-(η2-dppm)-8-(η1-dppm)-nido-8,7-RhSB9H10] and [9,9-(η2-dppm)-9-(η1-dppm)-nido-9,7,8-RhC2B8H11], Respectively. Inorg. Chem. 2002, 41, 5837–5843. [Google Scholar] [PubMed]
  58. Dou, J.; Wu, L.; Guo, Q.; Li, D.; Wang, D. A New Strategy for the Preparation of Metallaboranes-Solvothermal Synthesis and Structural Characterisation of Two Nido 11-vertex Diplatinaundecaborane Clusters. Eur. J. Inorg. Chem. 2005, 2005, 63–65. [Google Scholar] [CrossRef]
  59. Mingos, D.M.P. Polyhedral skeletal electron pair approach. Acc. Chem. Res. 1984, 17, 311–319. [Google Scholar] [CrossRef]
  60. Jung, C.W.; Hawthorne, M.F. Eleven-Vertex Rhodium, Iridium, and Ruthenium Phosphinometallocarborane Complexes Formed from Sodium Undecahydro-5,6-dicarba-nido-decaborate (1-). J. Am. Chem. Soc. 1980, 102, 3024–3032. [Google Scholar] [CrossRef]
  61. Ferguson, G.; Jennings, M.C.; Lough, A.J.; Coughlan, S.; Spalding, T.R.; Kennedy, J.D.; Fontaine, X.L.R.; Stibr, B. Novel Rhodathiaborane Complexes Derived from [(PPh3)2RhSB9H10]. J. Chem. Soc. Chem. Commun. 1990, 12, 891–894. [Google Scholar] [CrossRef]
  62. Coughlan, S.; Spalding, T.R.; Ferguson, G.; Gallagher, J.F.; Lough, A.J.; Fontaine, X.L.R.; Kennedy, J.D.; Stıbr, B. Metallaheteroborane chemistry. Part 10. Synthesis and characterisation of closo-structured rhodathiaborane complexes [1-(CO)-1-L-3-L′-1,2-RhSB9H8](L = L′ = PPh3; L = PMe2Ph, L′ = PMe2Ph or PPh3). J. Chem. Soc. Dalton Trans. 1992, 19, 2865–2871. [Google Scholar] [CrossRef]
  63. Volkov, O.; Macias, R.; Rath, N.P.; Barton, L. Chemistry on a metallathiaborane cluster Part 4: Reactions of 11-vertex rhodathiaboranes with bidentate phosphines and their subsequent rearrangements. J. Organomet. Chem. 2002, 657, 40–47. [Google Scholar] [CrossRef]
  64. Volkov, O.; Rath, N.P.; Barton, L. Chemistry of a Metallathiaborane Cluster. 5. Reaction of [8,8-(η2-dppm)-8-(η1-dppm)-nido-8,7-RhSB9H10] and Its Derivatives with Organotransition-Metal Reagents. Organometallics 2002, 21, 5505–5514. [Google Scholar] [CrossRef]
  65. Bould, J.; Cunchillos, C.; Lahoz, F.J.; Oro, L.A.; Kennedy, J.D.; Macias, R. New Iridathiaboranes with Reversible IsonidoNido Cluster Flexibility. Inorg. Chem. 2010, 49, 7353–7361. [Google Scholar] [CrossRef]
  66. Kasper, J.S.; Lucht, C.M.; Harker, D. The Crystal Structure of Decaborane, B10H14. Acta Cryst. 1950, 3, 436–455. [Google Scholar] [CrossRef]
  67. Kölle, U.; Khouzami, F.; Fuss, B. Bridged C5Me5CoII Complexes—Reactive Intermediates in the Cyclopentadienylation of Cobalt(II) Halides. Angew. Chem. Int. Ed. 1982, 21, 131–132. [Google Scholar] [CrossRef]
  68. Yoshino, T.; Ikemoto, H.; Matsunaga, S.; Kanai, M.A. Cationic High-Valent Cp*CoIII Complex for the Catalytic Generation of Nucleophilic Organometallic Species: Directed C–H Bond Activation. Angew. Chem. Int. Ed. 2013, 52, 2207–2211. [Google Scholar] [CrossRef] [PubMed]
  69. Lalancette, J.M.; Frêche, A.; Monteux, R. Reductions with sulfurated borohydrides. I. Preparation of sulfurated borohydrides. Can. J. Chem. 1968, 46, 2754–2757. [Google Scholar] [CrossRef]
  70. Lalancette, J.M.; Arnac, M. Reductions with sulfurated borohydrides. III. Borohydrides incorporating selenium and tellurium. Can. J. Chem. 1969, 47, 3695–3697. [Google Scholar] [CrossRef]
  71. Ryschkewitsch, G.E.; Nainan, K.C. Octahydrotriborate(1-) ([B3H8]-) salts. Method A. Inorg. Synth. 1974, 15, 113–114. [Google Scholar]
  72. Altomare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A.; Burla, M.C.; Polidori, G.; Camalli, M. SIRPOW. 92—A program for automatic solution of crystal structures by direct methods optimized for powder data. J. Appl. Cryst. 1994, 27, 435–436. [Google Scholar] [CrossRef]
  73. Sheldrick, G.M. SHELXT–Crystal structure refinement with SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar]
  74. Sheldrick, G.M. SHELXS97 and SHELXL97: Program for Crystal Structure Solution and Refinement; University of Gottingen: Gottingen, Germany, 1997. [Google Scholar]
  75. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  76. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.G.; Mennucci, B.; Petersson, G.A.; Nakatsuji, H.; Caricato, M.; et al. Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  77. Lee, C.; Yang, W.; Parr, R.G. Development of the Colic-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Wadt, W.R.; Hay, P.J. Ab Initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar] [CrossRef]
  79. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. A New Basis Set Exchange: An Open, Up-to-date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [Google Scholar] [CrossRef]
  80. London, F.J. Quantum theory of interatomic currents in aromatic combinations. J. Phys. Radium 1937, 8, 397–409. [Google Scholar] [CrossRef] [Green Version]
  81. Ditchfield, R. Self-consistent perturbation theory of diamagnetism. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  82. Wolinski, K.; Hinton, J.F.; Pulay, P. Efficient implementation of the gauge-independent atomic orbital method for NMR chemical shift calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  83. Chemcraft-Graphical Software for Visualization of Quantum Chemistry Computations. Available online: https://www.chemcraftprog.com (accessed on 21 March 2021).
  84. Glendening, E.D.; Badenhoop, J.K.; Reed, A.E.; Carpenter, J.E.; Bohmann, J.A.; Morales, C.M.; Landis, C.R.; Weinhold, F. NBO Program 6.0; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, USA, 2013. [Google Scholar]
  85. Wiberg, K. Application of the Pople-Santry-Segal CNDO Method to the Cyclopropylcarbinyl and Cyclobutyl Cation and to Bicyclobutane. Tetrahedron 1968, 24, 1083–1096. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of the open-faced tricobaltaborane cluster 1.
Scheme 1. Synthesis of the open-faced tricobaltaborane cluster 1.
Inorganics 09 00028 sch001
Figure 1. Molecular structure and atom labels of one of the residues of 1. Selected bond lengths (Å) and bond angles (°) of 1: B1–B2 1.82(2), B1–Co3 2.051(14), B1–Co2 2.063(16), B2–Co3 2.084(13), B6–B7 1.78(2), B4–Co1 2.067(14), B7–Co1 2.030(14), B7–B8′–B5 113.0(17), and B7–Co1–B4 87.4(6). Co–H–B bridging hydrogens could not be located.
Figure 1. Molecular structure and atom labels of one of the residues of 1. Selected bond lengths (Å) and bond angles (°) of 1: B1–B2 1.82(2), B1–Co3 2.051(14), B1–Co2 2.063(16), B2–Co3 2.084(13), B6–B7 1.78(2), B4–Co1 2.067(14), B7–Co1 2.030(14), B7–B8′–B5 113.0(17), and B7–Co1–B4 87.4(6). Co–H–B bridging hydrogens could not be located.
Inorganics 09 00028 g001
Figure 2. Schematic representation of stepwise generation of the cluster framework of 1 (B) from an icosahedron (A).
Figure 2. Schematic representation of stepwise generation of the cluster framework of 1 (B) from an icosahedron (A).
Inorganics 09 00028 g002
Chart 1. Examples of 11-vertex−12-SEP (skeletal electron pairs) nido-metalla(hetero)borane clusters.
Chart 1. Examples of 11-vertex−12-SEP (skeletal electron pairs) nido-metalla(hetero)borane clusters.
Inorganics 09 00028 ch001
Figure 3. Schematic representation of stepwise generation of the 11-vertex isonido (D) and 10-vertex nido (E) cluster frameworks from an octadecahedron (C) and their examples.
Figure 3. Schematic representation of stepwise generation of the 11-vertex isonido (D) and 10-vertex nido (E) cluster frameworks from an octadecahedron (C) and their examples.
Inorganics 09 00028 g003
Figure 4. Frontier molecular orbitals of 1’. Contour value: ±0.05 [e/bohr3]1/2.
Figure 4. Frontier molecular orbitals of 1’. Contour value: ±0.05 [e/bohr3]1/2.
Inorganics 09 00028 g004
Scheme 2. Synthesis of a symmetric dicobaltaborane analogue of decaborane(14).
Scheme 2. Synthesis of a symmetric dicobaltaborane analogue of decaborane(14).
Inorganics 09 00028 sch002
Figure 5. Molecular structure and atom labels for 2: frontal view (left) and bottom angular view (right, Cp* ligands attached to Co are omitted for clarity). Selected bond lengths (Å) and bond angles (°): B1–Co1 1.972(5), B2–B3 1.971(7), B2–Co1 2.043(5), B3–B4 1.807(8), B3–Co2 2.043(4), B5–Co2 2.041(5), B7–Co1 2.046(4), B3–B2–Co1 115.5(2), B4–B3–B2 116.3(4), B3–B4–B5 103.9(3), and B2–Co1–B7 89.1(2).
Figure 5. Molecular structure and atom labels for 2: frontal view (left) and bottom angular view (right, Cp* ligands attached to Co are omitted for clarity). Selected bond lengths (Å) and bond angles (°): B1–Co1 1.972(5), B2–B3 1.971(7), B2–Co1 2.043(5), B3–B4 1.807(8), B3–Co2 2.043(4), B5–Co2 2.041(5), B7–Co1 2.046(4), B3–B2–Co1 115.5(2), B4–B3–B2 116.3(4), B3–B4–B5 103.9(3), and B2–Co1–B7 89.1(2).
Inorganics 09 00028 g005
Figure 6. Frontier molecular orbital diagram of 2′, 3′, and 4′ (see text). Contour value: ±0.05 [e/bohr3]1/2.
Figure 6. Frontier molecular orbital diagram of 2′, 3′, and 4′ (see text). Contour value: ±0.05 [e/bohr3]1/2.
Inorganics 09 00028 g006
Table 1. Selected structural and spectroscopic parameters of 10-vertex−12-SEP nido-cobaltaborane clusters. a
Table 1. Selected structural and spectroscopic parameters of 10-vertex−12-SEP nido-cobaltaborane clusters. a
Geometryd(B-B)av [Å]11B NMR (δ,ppm)Ref.Geometryd(B-B)av [Å]11B NMR (δ,ppm)Ref.
Inorganics 09 00028 i001
6-[(Cp*Co)B9H13]
1.7820.5, 15.4, 5.2, −1.2, −12.4, −29.8[47] Inorganics 09 00028 i002
3,5-[(Cp*Co)2B8H12]
1.8042.4, 30.1, 25.3, 20.1, 9.1, 6.3, −1.2, −12.9[52]
Inorganics 09 00028 i003
5-[(Cp*Co)B9H13]
1.7829.4, 27.1, 12.5, 4.1, 3.5, 1.7, −1.1, −14.5, −36.9[52] Inorganics 09 00028 i004
6-Cl-5,7-[(Cp*Co)2B8H12]
-49.9, 23.1, 19.4, 7.1, −7.1, −40.7[47]
Inorganics 09 00028 i005
2-[(CpCo)B9H13]
-17.0, 14.6, 9.0, 1.7, −38.2[49] Inorganics 09 00028 i006
5,7-[(Cp*Co)2B8H12]
1.8057.3, 24.4, 6.0, −7.4, 40.3
[47]
Inorganics 09 00028 i007
[B10H14]
1.7811.7, 9.7, 0.28, −36.5[66] Inorganics 09 00028 i008 [(Cp*Co)2B8H11]1.7654.0, 29.7, 22.2, 8.7, 3.3, −40.1[51]
Inorganics 09 00028 i009
6,9-[(Cp*Co)2B8H12]
2.0220.8, 2.3[47] Inorganics 09 00028 i010
2,4-[(Cp*Co)2B8H12]
1.8126.6, 11.4, 5.6This wok
Inorganics 09 00028 i011
Table 2. Experimentally observed and calculated 1H and 11B chemical shifts for 24 species and 2′4′ models, respectively.
Table 2. Experimentally observed and calculated 1H and 11B chemical shifts for 24 species and 2′4′ models, respectively.
Cluster1H NMR δ (ppm)11B NMR δ (ppm)Ref.
Expt.Calc.Expt.Calc.
2/2′−4.25−6.2226.6, 11.4, 5.620.0, 11.0, 3.0This work
3/3′−3.52−5.4321.1, 9.4, 1.916.2, 6.5, 0.856 (expt.)
4/4″−3.73−5.451.8, −0.2, −12.97.3, −1.7, −11.027 (expt.)
Table 3. Crystallographic and structural refinement data for 1 and 2.
Table 3. Crystallographic and structural refinement data for 1 and 2.
Compound12
CCCDC No.20436792060786
Empirical formulaC30H53B8Co3C20H42B8Co2
Formula weight676.99486.87
Crystal systemTriclinicMonoclinic
Space groupP 1 ¯ P21/n
a (Å)8.785 (4)8.3902 (2)
b (Å)11.638 (6)40.6669 (12)
c (Å)16.993 (9)8.5097 (2)
α (°)96.244 (18)90
β (°)96.56 (2)119.0361 (11)
γ (°)101.532 (18)90
Volume (Å3)1675.6 (15)2538.60 (12)
Z24
ρcalc(g/cm−3)1.3421.274
μ (mm−1)1.4871.312
F (000)7081024
2θ range for data collection (°)3.6−44.92−50
Independent reflections 39904472
Final R indices [I >= 2σ (I)]R1 = 0.0889, wR2 = 0.2412R1 = 0.0439, wR2 = 0.0958
Parameters452329
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pathak, K.; Nandi, C.; Halet, J.-F.; Ghosh, S. Metal-Rich Metallaboranes: Synthesis, Structures and Bonding of Bi- and Trimetallic Open-Faced Cobaltaboranes. Inorganics 2021, 9, 28. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9040028

AMA Style

Pathak K, Nandi C, Halet J-F, Ghosh S. Metal-Rich Metallaboranes: Synthesis, Structures and Bonding of Bi- and Trimetallic Open-Faced Cobaltaboranes. Inorganics. 2021; 9(4):28. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9040028

Chicago/Turabian Style

Pathak, Kriti, Chandan Nandi, Jean-François Halet, and Sundargopal Ghosh. 2021. "Metal-Rich Metallaboranes: Synthesis, Structures and Bonding of Bi- and Trimetallic Open-Faced Cobaltaboranes" Inorganics 9, no. 4: 28. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics9040028

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

Article Metrics

Back to TopTop