Next Article in Journal
Macrocybin, a Natural Mushroom Triglyceride, Reduces Tumor Growth In Vitro and In Vivo through Caveolin-Mediated Interference with the Actin Cytoskeleton
Next Article in Special Issue
Synthesis and Structure of Nido-Carboranyl Azide and Its “Click” Reactions
Previous Article in Journal
The [DPPH●/DPPH-H]-HPLC-DAD Method on Tracking the Antioxidant Activity of Pure Antioxidants and Goutweed (Aegopodium podagraria L.) Hydroalcoholic Extracts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

The First Nickelacarborane with closo-nido Structure

by
Ekaterina P. Andreichuk
1,2,
Sergey A. Anufriev
1,
Kyrill Yu. Suponitsky
1,3 and
Igor B. Sivaev
1,*
1
A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilov Street, 119991 Moscow, Russia
2
Higher Chemical College at the Russian Academy of Sciences, D.I. Mendeleev Russian Chemical Technological University, 9 Miusskaya Square, 125047 Moscow, Russia
3
N.S. Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, 31 Leninskii Avenue, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(24), 6009; https://doi.org/10.3390/molecules25246009
Submission received: 25 November 2020 / Revised: 16 December 2020 / Accepted: 18 December 2020 / Published: 18 December 2020

Abstract

:
The first nickelacarborane with closo-nido structure [10′,11′-(Py)2-3,9′-Ni(1,2-C2B9H11)(7′,8′-C2B8H8)] was isolated from the reaction of nickel(IV) bis(dicarbollide) with pyridine. The molecular structure of this complex was determined by single crystal X-ray diffraction. The nickel atom is a common vertex for the closo-NiC2B9 cluster and the nido-NC2B8 cluster where it is located together with carbon atoms in the open NiC2B2 pentagonal face. It is assumed that its formation proceeds through the nucleophile-induced removal of the B(6)H vertex followed by rearrangement of the forming 11-vertex cluster, which most likely proceeds through a sequence of closing and opening reactions.

1. Introduction

Due to the practically unlimited possibility of varying the structure of carboranes, including by replacing hydrogen atoms with various groups, transition metal complexes with carborane ligands are one of the most exciting areas of modern chemistry [1,2,3]. In addition to purely academic interest, a number of them show excellent prospects for practical use. Thus, nickel bis(dicarbollide) and its derivatives [4] attract interest of researchers due to their potential applications in materials science, including molecular switches [5,6,7,8], solar cells [9,10] and conductive metal-organic frameworks [11]. At the same time, the available information on the stability of these complexes is fragmentary and rather contradictory [12,13,14,15]. This prompted us to study the stability of nickel(IV) bis(dicarbollide) under various conditions. Earlier, we found that boiling nickel bis(dicarbollide) in ethanol leads to its decomposition with the formation of a mixture of nido-carborane and its 3-ethoxy derivative as the main products [16].
In this contribution we report the formation of the first nickelacarborane with a closo-nido structure in the reaction of nickel(IV) bis(dicarbollide) with boiling pyridine.

2. Results and Discussion

It was found that prolonged refluxing nickel(IV) bis(dicarbollide) (1) in pyridine leads to its complete transformation into a mixture of various nickelacarboranes, mainly of paramagnetic nature. The separation of this mixture by column chromatography on silica gave a diamagnetic nickelacarborane [10′,11′-(Py)2-3,9′-Ni(1,2-C2B9H11)(7′,8′-C2B8H8)] (2), the structure of which was established by single crystal X-ray diffraction. The reaction is characterized by good reproducibility with an isolated yield of 2 ranging from 19 to 23% (Scheme 1).
The 1H NMR spectrum of 2 contains a set of signals from two different N-substituted pyridines in the range of 8.9–7.6 ppm and four signals from different CH groups of carborane ligands at 3.55, 3.33, 2.80 and 1.87 ppm, while the 11B NMR spectrum contains two sets of signals corresponding to symmetrical and asymmetrical carborane ligands (See Supplementary Materials). It should be noted that determination of the structure of metallacarboranes containing carborane ligands of different geometry is a difficult task that normally cannot be solved using methods of NMR spectroscopy due to signal overlap and, therefore, requires single crystal X-ray diffraction study [17,18,19,20,21,22].
An asymmetric unit cell of the complex 2 contains one molecule (Figure 1). In the structure of the complex, the nickel atom is a common vertex for the closo-NiC2B9 cluster and the nido-NC2B8 cluster, where it is located together with carbon atoms in the open NiC2B2 pentagonal face. Each of the remaining two boron atoms on the pentagonal face is bonded to a pyridine molecule.
The Ni-B bonds in the nido-fragment (2.093(3)–2.119(3) Å) are only slightly shorter than in the closo-fragment (2.103(2)–2.127(3) Å), while the Ni-C bond lengths differ more significantly (2.002(2) and 2.105(2)–2.170(2) Å for the nido- and closo-fragments, respectively). The same trend is observed also for the C-C bond lengths (1.508(3) and 1.573(3) Å for the nido- and closo-fragments, respectively). The B-N lengths equal 1.553(3) and 1.561(3) Å, that are somewhat shorter than in a related cobaltacarborane (Et4N)[11′-Py-3,9′-Co(1,2-C2B9H11)(7′,8′-C2B8H10)] (1.592(11) Å) [23], but close to those found in pyridinium derivatives of nido-carborane [9-Py-7,8-C2B9H11] (1.546(2) Å) [24] and metallacarboranes [4-Py-3-(C4Me4)-3,1,2-CoC2B9H10] (1.548(3) Å) [24], [8-Py-3,3′-Co(1,2-C2B9H10)(1′,2′-C2B9H11)] (1.556(5) Å) [25], [8-Py-3,3′-Fe(1,2-C2B9H10)(1′,2′-C2B9H11)] (1.548(6) and 1.553(6) Å) [25]. The relative orientation of the carborane ligands can be described by the pseudotorsion angle C1 Centroid (C1-C2-B7-B8-B4) Ni1-B10’ that is equal to 104.0 (2)°. Slightly shortened intramolecular CH⋯HB contacts were found between the pyridine substituent closest to the metal atom and the dicarbollide ligands. To assess their possible contribution to the stabilization of the observed conformation of complex 2, we carried out its QTAIM (quantum theory of atoms in molecules) study [26,27]. Energy of intramolecular noncovalent interactions was estimated based on their correlation with the energy density function [28] that was found to be reliable for different types of weak noncovalent interactions [29,30,31]. Two CH⋯HB noncovalent contacts (shown in Figure 1) were observed (B5’-H5’⋯H3A-C3, 2.40Å, −2.4 kcal/mol; B8-H8⋯H7A-C7, 2.31Å, −2.6 kcal/mol).
It should be noted that neither other transition metal bis(dicarbollide) complexes, nor ortho-carborane itself, do not undergo deboronation with pyridine, whereas C-halogen derivatives were found to be accessible for nucleophilic attack of pyridine and its derivatives [32]. Complex 2 is the first example of a nickelacarborane in which the metal atom is common to both the closo-icosahedral cluster and the nido-cluster formed by removing one vertex from the icosahedron. A related cobaltacarborane [11′-Py-3,9′-Co(1,2-C2B9H11)(7′,8′-C2B8H10)] (3) was earlier prepared by removing one BH vertex on heating the parent cobalt bis(dicarbollide) [3,3′-Co(1,2-C2B9H11)2] in a 25% aqueous solution of potassium hydroxide at 95 °C, followed by the oxidation of the formed nido- structure to the closo- one with 30% hydrogen peroxide, and the reaction with pyridine, which leads to the opening of the 11-vertex metallacarborane the substitution of one hydrogen atom with pyridine [33,34]. In general, the formation of complex 2 can be described as a polyhedral contraction reaction [33,35,36].
Taking into account that the BH group at position 6 is the most susceptible to nucleophilic attack in the 3,1,2-MC2B9 metallacarboranes, we believe the formation of complex 2 should also proceed through the removal of this vertex followed by rearrangement of the 11-vertex cluster, which most likely proceeds through a sequence of closing and opening reactions, rather than via detachment and reattachment of the carborane ligand. Therefore, it is reasonable to assume that other formed nickelacarboranes are related to cobaltacarboranes, which are intermediate products of the formation of complex 3. The identification of these products is in progress.

3. Materials and Methods

3.1. General Methods

Nickel(IV) bis(dicarbollide) (1) was prepared according to the published procedure [37]. The reaction progress was monitored by thin layer chromatography (Merck F254 silica gel on aluminum plates) and visualized using 0.5% PdCl2 in 1% HCl in aq. MeOH (1:10). Acros Organics silica gel (0.060–0.200 mm) was used for column chromatography. The NMR spectra at 400 MHz (1H), 128 MHz (11B) and 100 MHz (13C) in acetone-d6 were recorded with a Varian Inova 400 spectrometer. The residual signal of the NMR solvent relative to tetramethylsilane was taken as the internal reference for 1H and 13C NMR spectra. 11B NMR spectra were referenced using BF3·Et2O as an external standard.
The X-ray diffraction experiment for compound 2 was carried out using a SMART APEX2 CCD diffractometer (λ(Mo-Kα) = 0.71073 Å, graphite monochromator, ω-scans) at 120 K. Collected data were processed by the SAINT and SADABS programs incorporated into the APEX2 program package [38]. The structure was solved by direct methods and refined by the full-matrix least-squares procedure against F2 in anisotropic approximation. The refinement was carried out with the SHELXTL program [39]. The CCDC number (2046492) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif.

3.2. Reaction of Nickel(IV) Bis(Dicarbollide) with Pyridine

A solution of nickel(IV) bis(dicarbollide) (50 mg, 0.3 mmol) in 30 mL of pyridine was stirred under reflux for 50 h. Thereafter, solution was cooled to room temperature and concentrated under reduced pressure. The crude product was subjected to column chromatography on silica using of dichloromethane as eluent. The third boron-containing fraction gave dark crystals of 2 (17 mg, yield 23%). 1H NMR: δ 8.89 (d, 2H, J = 7.0 Hz, o-CHPy), 8.70 (2H, d, J = 6.8 Hz, o-CHPy), 8.46 (1H, t, J = 7.0 Hz, p-CHPy), 8.19 (1H, t, J = 6.8 Hz, p-CHPy), 8.01 (2H, t, J = 7.0 Hz, m-CHPy), 7.60 (2H, t, J = 6.8 Hz, m-CHPy), 4.0–0.4 (15H, br, BHCarb), 3.55 (1H, s, CHCarb), 3.33 (1H, s, CHCarb), 2.80 (1H, s, CHCarb), 1.87 (1H, s, CHCarb) ppm; 11B NMR: δ 15.2 (1B, s, B-N), 7.7 (1B, d, J = 116 Hz), 2.3 (1B, d, J = 149 Hz), 0.8 (1B, s, B-N), –2.1 (1B, d, J = 146 Hz), –4.6 (2B, d, J = 137 Hz), –6.4 (2B, d, J = 146 Hz), –7.8 (1B, d, J = 114 Hz), –10.9 (1B, d, J = 133 Hz,), –17.4 (3B, d, J = 138 Hz), –20.2 (1B, d, J = 163 Hz), –22.9 (1B, d, J = 163 Hz), –31.6 (1B, d, J = 139 Hz) ppm; 13C NMR: δ 149.8 (o-CHPy), 146.6 (o-CHPy), 143.5 (p-CHPy), 143.1 (p-CHPy), 128.6 (m-CHPy), 126.1 (m-CHPy), 54.1 (CHCarb), 45.7 (CHCarb), 40.9 (CHCarb), 40.2 (CHCarb) ppm. Crystallographic data: C14H29B17N2Ni are monoclinic, space group P21/n: a = 13.0440(8) Å, b = 9.8669(6) Å, c = 18.6798(11) Å, β= 98.2660(10)°, V = 2379.2(2) Å3, Z = 4, M = 467.87, dcryst = 1.306 g·cm−3. wR2 = 0.0990 calculated on F2hkl for all 6244 independent reflections with 2θ < 56.4°, (GOF = 1.020, R = 0.0408 calculated on Fhkl for 4310 reflections with I > 2σ(I)).

Supplementary Materials

The following are available online, NMR spectra of compound 2.

Author Contributions

Synthesis, E.P.A.; NMR spectroscopy study, S.A.A.; single crystal X-ray diffraction, K.Y.S.; supervision and manuscript concept, I.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (Grant No. 16-13-10331).

Acknowledgments

The NMR spectroscopy and X-ray diffraction data were obtained by using equipment from the Center for Molecular Structure Studies at A.N. Nesmeyanov Institute of Organoelement Compounds, operating with support from the Ministry of Science and Higher Education of the Russian Federation.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Grimes, R.N. Transition metal metallacarbaboranes. In Comprehensive Organometallic Chemistry II, Volume 1; Housecraft, C.E., Ed.; Pergamon Press: Oxford, UK, 1995; pp. 373–430. [Google Scholar] [CrossRef]
  2. Hosmane, N.S.; Maguire, J.A. Metallacarboranes of d- and ƒ-block metals. In Comprehensive Organometallic Chemistry III, Volume 3; Housecraft, C.E., Ed.; Elsevier Science: Amsterdam, The Netherlands, 2007; pp. 175–264. [Google Scholar] [CrossRef]
  3. Grimes, R.N. Carboranes, 3rd ed.; Academic Press: London, UK, 2016; pp. 711–903. [Google Scholar] [CrossRef]
  4. Sivaev, I.B.; Bregadze, V.I. Chemistry of nickel and iron bis(dicarbollides). A review. J. Organomet. Chem. 2000, 614–615, 27–36. [Google Scholar] [CrossRef]
  5. Hawthorne, M.F.; Zink, J.I.; Skelton, J.M.; Bayer, M.J.; Liu, C.; Livshits, E.; Baer, R.; Neuhauser, D. Electrical or photocontrol of the rotary motion of a metallacarborane. Science 2004, 303, 1849–1851. [Google Scholar] [CrossRef] [Green Version]
  6. Safronov, A.V.; Shlyakhtina, N.I.; Everett, T.A.; VanGordon, M.R.; Sevryugina, Y.V.; Jalisatgi, S.S.; Hawthorne, M.F. Direct observation of bis(dicarbollyl)nickel conformers in solution by fluorescence spectroscopy: An approach to redox-controlled metallacarborane molecular motors. Inorg. Chem. 2014, 53, 10045–10053. [Google Scholar] [CrossRef] [PubMed]
  7. Shlyakhtina, N.I.; Safronov, A.V.; Sevryugina, Y.V.; Jalisatgi, S.S.; Hawthorne, M.F. Synthesis, characterization, and preliminary fluorescence study of a mixed-ligand bis(dicarbollyl)nickel complex bearing a tryptophan-BODIPY FRET couple. J. Organomet. Chem. 2015, 798, 234–244. [Google Scholar] [CrossRef]
  8. Anufriev, S.A.; Suponitsky, K.Y.; Filippov, O.A.; Sivaev, I.B. Synthesis and structure of methylsulfanyl derivatives of nickel bis(dicarbollide). Molecules 2019, 24, 4449. [Google Scholar] [CrossRef] [Green Version]
  9. Spokoyny, A.M.; Li, T.C.; Farha, O.K.; Machan, C.W.; She, C.; Stern, C.L.; Marks, T.J.; Hupp, J.T.; Mirkin, C.A. Electronic tuning of nickel-based bis(dicarbollide) redox shuttles in dye-sensitized solar cells. Angew. Chem. Int. Ed. 2010, 49, 5339–5343. [Google Scholar] [CrossRef]
  10. Li, T.C.; Spokoyny, A.M.; She, C.; Farha, O.K.; Mirkin, C.A.; Marks, T.J.; Hupp, J.T. Ni(III)/(IV) bis(dicarbollide) as a fast, noncorrosive redox shuttle for dye-sensitized solar cells. J. Am. Chem. Soc. 2010, 132, 4580–4582. [Google Scholar] [CrossRef]
  11. Kung, C.-W.; Otake, K.; Buru, C.T.; Goswami, S.; Cui, Y.; Hupp, J.T.; Spokoyny, A.M.; Farha, O.K. Increased electrical conductivity in a mesoporous metal-organic framework featuring metallacarboranes guests. J. Am. Chem. Soc. 2018, 140, 3871–3875. [Google Scholar] [CrossRef]
  12. Warren, L.F.; Hawthorne, M.F. Chemistry of the bis[π-(3)-1,2-dicarbollyl] metalates of nickel and palladium. J. Am. Chem. Soc. 1970, 92, 1157–1173. [Google Scholar] [CrossRef]
  13. Maier, N.A.; Erdman, A.A.; Zubreichuk, Z.P.; Prokopovich, V.P.; Ol’dekop, Y.A. Synthesis and some transformations of complex nickel(II) salts of bis(3,1,2-dicarbollyl)nickel(III). Preparation of 3-(2,2′- bipyridyl)-closo-3,1,2-nickeladicarbadodecaborane. J. Organomet. Chem. 1985, 292, 297–302. [Google Scholar] [CrossRef]
  14. Zubreichuk, Z.P.; Erdman, A.A.; Ivko, A.A.; Maier, N.A. Synthesis and some transformations of complex salts of bis-o-dicarbollyliron(II), -cobalt(II), and -nickel(II). Russ. J. Gen. Chem. 2001, 71, 531–534. [Google Scholar] [CrossRef]
  15. Erdman, A.A.; Zubreichuk, Z.P.; Knizhnikov, V.A.; Maier, A.A.; Aleksandrov, G.G.; Nefedov, S.E.; Eremenko, I.L. Synthesis and the structure of the triphenylphosphine complex of o-nickelacarborane, 3,3-(PPh3)2-3,1,2-NiC2B9H11. Russ. Chem. Bull. 2001, 50, 2248–2250. [Google Scholar] [CrossRef]
  16. Andreichuk, E.P.; Anisimov, A.A.; Shmalko, A.V.; Suponitsky, K.Y.; Sivaev, I.B.; Bregadze, V.I. Stability of nickel bis(dicarbollide) complexes. Mendeleev Commun. 2019, 29, 534–536. [Google Scholar] [CrossRef]
  17. Viñas, C.; Pedrajas, J.; Teixidor, T.; Kivekäs, R.; Sillanpää, R.; Welch, A.J. First example of a bis(dicarbollide) metallacarborane containing a B,C‘-heteronuclear bridge. Inorg. Chem. 1997, 36, 2988–2991. [Google Scholar] [CrossRef] [PubMed]
  18. Planas, J.G.; Viñas, C.; Teixidor, F.; Light, M.E.; Hursthouse, M.B. A boron-boron linked large metallacarborane cluster: Characterization and X-ray structure of 8,9′-[closo-{3-Co(η5-C5H5)-1,2-C2B9H10}]2. J. Organomet. Chem. 2006, 691, 3472–3476. [Google Scholar] [CrossRef]
  19. Molotkov, A.P.; Vinogradov, M.M.; Moskovets, A.P.; Chusova, O.; Timofeev, S.V.; Fastovskiy, V.A.; Nelyubina, Y.V.; Pavlov, A.A.; Chusov, D.A.; Loginov, D.A. Iridium halide complexes [1,1-X2-8-SMe2-1,2,8-IrC2B9H10]2 (X = Cl, Br, I): Synthesis, reactivity and catalytic activity. Eur. J. Inorg. Chem. 2017, 4635–4644. [Google Scholar] [CrossRef] [Green Version]
  20. Jeans, R.J.; Chan, A.P.Y.; Riley, L.E.; Taylor, J.; Rosair, G.M.; Welch, A.J.; Sivaev, I.B. Arene-ruthenium complexes of 1,1′-bis(ortho-carborane): Synthesis, characterization, and catalysis. Inorg. Chem. 2019, 58, 11751–11761. [Google Scholar] [CrossRef]
  21. Chan, A.P.Y.; Parkinson, J.A.; Rosair, G.M.; Welch, A.J. Bis(phosphine)hydridorhodacarborane derivatives of 1,1′-bis(ortho-carborane) and their catalysis of alkene isomerization and the hydrosilylation of acetophenone. Inorg. Chem. 2020, 59, 2011–2023. [Google Scholar] [CrossRef]
  22. Chan, A.P.Y.; Rosair, G.M.; Welch, A.J. Exopolyhedral ligand orientation controls diastereoisomer in mixed-metal bis(carboranes). Molecules 2020, 25, 519. [Google Scholar] [CrossRef] [Green Version]
  23. Churchill, M.R.; Gold, K. Geometry of the (B8C2H10·C5H5N2−) anion from an X-ray structural analysis of [Et4N+][(B9C2H11)Co(B8C2H10·C5H5N)]. J. Chem. Soc. Chem. Commun. 1972, 901–902. [Google Scholar] [CrossRef]
  24. Meshcheryakov, V.I.; Kitaev, P.S.; Lyssenko, K.A.; Starikova, Z.A.; Petrovskii, P.V.; Janoušek, Z.; Corsini, M.; Laschi, F.; Zanello, P.; Kudinov, A.R. (Tetramethylcyclobutadiene)cobalt complexes with monoanionic carborane ligands [9-L-7,8-C2B9H10] (L = SMe2, NMe3 and py). J. Organomet. Chem. 2005, 690, 4745–4754. [Google Scholar] [CrossRef]
  25. Buades, A.B.; Arderiu, V.S.; Olid-Britos, D.; Viñas, C.; Sillanpää, R.; Haukka, M.; Fontrodona, X.; Paradinas, M.; Ocal, C.; Teixidor, F. Electron accumulative molecules. J. Am. Chem. Soc. 2018, 140, 2957–2970. [Google Scholar] [CrossRef] [PubMed]
  26. Bader, R.F.W. Atoms in Molecules. A Quantum Theory; Clarendon Press: Oxford, UK, 1990. [Google Scholar]
  27. Keith, T.A. AIMAll (Version 15.05.18); TK Gristmill Software: Overland Park, KS, USA, 2015. [Google Scholar]
  28. Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen bond strengths revealed by topological analyses of experimentally observed electron densities. Chem. Phys. Lett. 1998, 285, 170–173. [Google Scholar] [CrossRef]
  29. Anufriev, S.A.; Sivaev, I.B.; Suponitsky, K.Y.; Godovikov, I.A.; Bregadze, V.I. Synthesis of 10-methylsulfide and 10-alkylmethylsulfonium nido-carborane derivatives: B-H⋯π Interactions between the B-H-B hydrogen atom and alkyne group in 10-RC≡CCH2S(Me)-7,8-C2B9H11. Eur. J. Inorg. Chem. 2017, 4436–4443. [Google Scholar] [CrossRef] [Green Version]
  30. Dmitrienko, A.O.; Karnoukhova, V.A.; Potemkin, A.A.; Struchkova, M.I.; Kryazhevskikh, I.A.; Suponitsky, K.Y. The influence of halogen type on structural features of compounds containing α-halo-α,α-dinitroethyl moieties. Chem. Heterocycl. Comp. 2017, 53, 532–539. [Google Scholar] [CrossRef]
  31. Sheremetev, A.B.; Aleksandrova, N.S.; Semyakin, S.S.; Suponitsky, K.Y.; Lempert, D.B. Synthesis and characterization of 3-(5-(Fluorodinitromethyl)-1H-1,2,4-triazol-3-yl)-4-nitrofurazan: A novel promising energetic component of boron-based fuels for rocket ramjet engines. Chem. Asian J. 2019, 14, 4255–4261. [Google Scholar] [CrossRef] [PubMed]
  32. Taoda, Y.; Sawabe, T.; Endo, Y.; Yamaguchi, K.; Fujii, S.; Kagechika, H. Identification of an intermediate in the deboronation of ortho-carborane: An adduct of ortho-carborane with two nucleophiles on one boron atom. Chem. Commun. 2008, 2049–2051. [Google Scholar] [CrossRef]
  33. Jones, C.J.; Francis, J.N.; Hawthorne, M.F. New 10- and 11-atom polyhedral metallocarboranes prepared by polyhedral contraction. J. Am. Chem. Soc. 1972, 94, 8391–8399. [Google Scholar] [CrossRef]
  34. Jones, C.J.; Francis, J.N.; Hawthorne, M.F. Derivative chemistry of metallocarboranes. nido 11-Atom metallocarboranes and their Lewis base adducts. J. Am. Chem. Soc. 1973, 95, 7633–7643. [Google Scholar] [CrossRef]
  35. Pisareva, I.V.; Dolgushin, F.M.; Tok, O.L.; Konoplev, V.E.; Suponitsky, K.Y.; Yanovsky, A.I.; Chizhevsky, I.T. Small metallacarborane closo-C2B4H6Ru(PPh3)2HCl formed from nido-5,6-C2B8H12 by ruthenium- mediated polyhedral contraction. Organometallics 2001, 20, 4216–4220. [Google Scholar] [CrossRef]
  36. Konoplev, V.E.; Tachaev, M.V.; Ulyukina, E.A. “Polyhedral contraction” of tetramethylammonium closo-carboundecaborate underaction of tris(triphenylphosphine)ruthenium dichloride. Izv. Vyssh. Uchebn. Zaved. Khim. Khim. Tekhnol. 2020, 63, 26–32. [Google Scholar] [CrossRef]
  37. Hawthorne, M.F.; Young, D.C.; Andrews, T.D.; Howe, D.V.; Pilling, R.L.; Pitts, A.D.; Reintjes, M.; Warren, L.F.; Wegner, P.A. π-Dicarbollyl derivatives of the transition metals. Metallocene analogs. J. Am. Chem. Soc. 1968, 90, 879–896. [Google Scholar] [CrossRef]
  38. APEX2 and SAINT; Bruker AXS Inc.: Madison, WI, USA, 2014.
  39. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Reaction of nickel(IV) bis(dicarbollide) with pyridine.
Scheme 1. Reaction of nickel(IV) bis(dicarbollide) with pyridine.
Molecules 25 06009 sch001
Figure 1. General view of complex 2 showing atomic numbering intramolecular CH⋯HB contacts. Thermal ellipsoids are drawn at 50% probability level.
Figure 1. General view of complex 2 showing atomic numbering intramolecular CH⋯HB contacts. Thermal ellipsoids are drawn at 50% probability level.
Molecules 25 06009 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Andreichuk, E.P.; Anufriev, S.A.; Suponitsky, K.Y.; Sivaev, I.B. The First Nickelacarborane with closo-nido Structure. Molecules 2020, 25, 6009. https://doi.org/10.3390/molecules25246009

AMA Style

Andreichuk EP, Anufriev SA, Suponitsky KY, Sivaev IB. The First Nickelacarborane with closo-nido Structure. Molecules. 2020; 25(24):6009. https://doi.org/10.3390/molecules25246009

Chicago/Turabian Style

Andreichuk, Ekaterina P., Sergey A. Anufriev, Kyrill Yu. Suponitsky, and Igor B. Sivaev. 2020. "The First Nickelacarborane with closo-nido Structure" Molecules 25, no. 24: 6009. https://doi.org/10.3390/molecules25246009

Article Metrics

Back to TopTop