Next Article in Journal
Synthesis and Biological Evaluation of Novel 2-Arylalkylthio-5-iodine-6-substituted-benzyl-pyrimidine-4(3H)-ones as Potent HIV-1 Non-Nucleoside Reverse Transcriptase Inhibitors
Next Article in Special Issue
Ruthenium Complexes as NO Donors for Vascular Relaxation Induction
Previous Article in Journal
Synthesis and Cytotoxicity of 2,3-Enopyranosyl C-Linked Conjugates of Genistein
Previous Article in Special Issue
Experimental and Theoretical Perspectives of the Noyori-Ikariya Asymmetric Transfer Hydrogenation of Imines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Unusual Conversion of Paramagnetic [3-Cl-3,3,8-{Ph2P(CH2)nPPh-µ-(C6H4-ortho)}-1,2-(CH3)2-closo-3,1,2-RuIIIC2B9H8] (n = 3 and 4) to Form the First 18-Electron P-Phenylene ortho-Cycloboronated closo-Ruthenacarboranes with a Dioxygen Ligand

by
Alexander Y. Kostukovich
,
Dmitrii I. D'yachihin
,
Fedor M. Dolgushin
,
Alexander F. Smol'yakov
,
Ivan A. Godovikov
and
Igor T. Chizhevsky
*
A.N. Nesmeyanov Institute of Organoelement Compounds of the RAS, 28 Vavilov Street, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(6), 7094-7103; https://doi.org/10.3390/molecules19067094
Submission received: 21 April 2014 / Revised: 16 May 2014 / Accepted: 22 May 2014 / Published: 30 May 2014
(This article belongs to the Special Issue Ruthenium Complex)

Graphical Abstract

1. Introduction

Boron-containing transition metal clusters exhibiting ortho-cycloboronation of the phosphine or diphosphine ligands on a metal atom are long known in metallaborane [1,2,3] and monocarbon metallacarborane [4,5] chemistry. However, first examples of related 12-vertex dicarbon metallacarborane complexes were reported only recently [6,7,8,9,10,11], and all known complexes of these series are RuIII species having a 17-electron shell. In this regard it is of interest that these paramagnetic ortho-cycloboronated metallacarborane systems demonstrate the exceptionally high efficiency as catalyst precursors for the atom transfer radical polymerization (ATRP) of methyl methacrylate, as well as of other vinyl monomers [7,8,9,10,11,12,13,14]. On the other hand, such electron-deficient complexes hold considerable promise in a variety of ways for the development of ruthenacarborane chemistry. We now report an initial study on the reactivity of paramagnetic ortho-phenylenecycloboronated closo-ruthenacarboranes resulting in preparation of a short series of new dioxygen ruthenium-carborane complexes.

2. Results and Discussion

Polyhedral contraction reactions of transition metals metallacarboranes occurring under base-mediated conditions to produce smaller metallacarborane species have not been extensively reported [15,16,17,18]. Hence, we proposed to extend studies in this area starting from the previously untested ortho-cycloboronated P-phenylene paramagnetic species [3-Cl-3,3,8-[Ph2P(CH2)nPPh-µ-( Molecules 19 07094 i002H8] (1,2-dimethyl-3-chloro-3,8-µ-{ortho-[phenyl-(3'-diphenylphosphinopropyl-κP(Ru))phosphino]phenyl-P(Ru),C1(B8)}-1,2-dicarba-3-ruthena-closo-dodecaborane (1) n = 3 [7] and 1,2-dimethyl-3-chloro-3,8-μ-{ortho-[phenyl-(4'-diphenylphosphinobutyl-κP(Ru))phosphino]phenyl-P(Ru),C1(B8)}-1,2-dicarba-3-ruthena-closo-dodecaborane (2); n = 4 [9]) with the hope of preparing either ortho-cycloboronated 11-vertex {RuC2B8} complexes or smaller clusters of these series. Accordingly, we examined the reaction of 1 and 2 with a fourfold molar excess of KOH at room temperature in a 1:1 benzene/methanol mixture in air.
Only ruthenacarborane products were isolated from these reactions, each in ca. 45% yield, after column chromatography on silica gel. These species proved, however, to exhibit different structures from those anticipated. It was evident from crystallographic data that the isolated species are actually novel 18-electron diamagnetic closo-ruthenacarboranes with the dioxygen ligand at the metal vertex, viz., [3-(η2-O2)-3,3,8-[Ph2P(CH2)nPPh-µ-( Molecules 19 07094 i003H8] (3, n = 3; 4, n = 4), respectively (Scheme 1).
Scheme 1. Schematic representation of the reactions of 1 and 2 with an oxygen.
Scheme 1. Schematic representation of the reactions of 1 and 2 with an oxygen.
Molecules 19 07094 g003
Single-crystal X-ray diffraction studies were carried out on both species 3 and 4. Their selected bond distances and angles are summarized in Table 1. Figure 1, on the top, shows the molecular structure of complex 3. In the crystal of complex 4, there are two independent molecules, which show only small differences in their geometry (see Table 1), and the molecular structure of only one of them is therefore shown in Figure 1, on the bottom.
Table 1. Selected bond lengths (Å) and angles (deg.) in 3 and 4 (two independent molecules A and B).
Table 1. Selected bond lengths (Å) and angles (deg.) in 3 and 4 (two independent molecules A and B).
Compound34
AB
Ru(3)-O(1)2.037(2)2.036(3)2.015(3)
Ru(3)-O(2)2.040(2)2.041(3)2.043(3)
Ru(3)-P(1)2.3412(7)2.3580(13)2.3693(13)
Ru(3)-P(2)2.2928(7)2.3201(13)2.3280(13)
Ru(3)-C(1)2.320(3)2.316(5)2.318(5)
Ru(3)-C(2)2.336(3)2.356(5)2.338(4)
Ru(3)-B(4)2.251(3)2.255(5)2.282(5)
Ru(3)-B(7)2.251(3)2.258(5)2.290(6)
Ru(3)-B(8)2.253(3)2.259(5)2.274(5)
O(1)-O(2)1.403(3)1.399(4)1.404(4)
C(1)-C(2)1.621(4)1.633(7)1.640(6)
O(1)-Ru(3)-O(2)40.26(8)40.14(12)40.49(12)
O(1)-Ru(3)-P(2)109.21(6)112.24(10)115.91(10)
O(2)-Ru(3)-P(2)79.80(6)81.07(10)83.70(9)
O(1)-Ru(3)-P(1)75.32(6)74.19(9)74.10(9)
O(2)-Ru(3)-P(1)100.35(6)98.09(10)95.59(10)
P(2)-Ru(3)-P(1)86.76(3)89.17(4)86.97(5)
O(2)-O(1)-Ru(3)69.97(11)69.76(18)68.68(18)
O(1)-O(2)-Ru(3)69.77(11)70.10(18)70.82(18)
According to the crystallographic data, the ruthenium atom in both 3 and 4 is bound in a η5-fashion to the C2B3 open face of C,C'-dimethylated nido-{C2B9}-carborane cage ligands, which have ortho-phenylenecycloboronated linkages connecting one of the P-phenyl rings of the ruthenium-bound dppp or dppb ligand, respectively, to the cluster B(8) atom. As expected, the presence of ortho-cycloboronated bridging fragment in both 3 and 4 accounts for the observed differences in the Ru-P(1) and Ru-P(2) bond lengths (2.3412(7) and 2.2928(7) Å in 3; av. 2.3637(13) and 2.3241(13) Å in 4), respectively. In addition, the metal atom in both structures 3 and 4 is coordinated by the dioxygen ligand in almost symmetrical η2-fashion with Ru-O distances range 2.015(3)-2.043(3) Å (av. 2.035 Å). The O-O distance in 3 (1.403(3) Å) and in 4 (1.399(4), 1.404(4) Å) is comparable to those found in the known oxygen Cp* ruthenium complexes such as [Cp*Ru(O2)(dppe)]PF6 (1.398(5)) Å [19]) or [Cp*Ru(O2)(dppm)]BPh4 (1.37(1)) Å [20], however, is somewhat shorter as compared with that in the only structurally studied oxygen iridium-carborane species with the exo-polyhedrally bound metal atom, [(Ph3P)2IrO2(7,8-(MeS)2-7,8-C2B9H10)]•Me2CO (1.471(8) Å) [21]. The carborane C-C bond distances in 3 (1.621(4) Å) and 4 (1.633(7) and 1.640(6) Å), as well as the other cluster C-B (1.69–1.75 Å) and B-B (1.73–1.84 Å) connectivities, are typical of 12-vertex closo-metallacarboranes [22].
Figure 1. ORTEP representation of the molecular structures of complexes 3 (top) and 4 (bottom) with thermal ellipsoids at the 50% probability level. Hydrogen atoms are omitted for clarity.
Figure 1. ORTEP representation of the molecular structures of complexes 3 (top) and 4 (bottom) with thermal ellipsoids at the 50% probability level. Hydrogen atoms are omitted for clarity.
Molecules 19 07094 g001
Analytical, IR and NMR spectroscopic data are entirely consistent with the formulation of complexes 3 and 4 made based on the above X-ray diffraction experiments. The Nujol mull infrared spectra of 3 and 4 show a broad band at ca. 940 cm−1. Since no such bands were observed in the IR spectra of starting complexes 1 and 2, these bands in IR spectra of 3 and 4 were assigned to νO-O. The detailed examination of the 1H, 13C{1H}, and 31P{1H} NMR spectra of complexes 3 and 4 provided additional useful information on their structures. The 31P{1H} NMR spectra of complexes 3 and 4 each contains two resonances: two sharp doublets (δP(1) = 18.6 ppm, δP(2) = 52.9; JP(1)-P(2) = 75.3 Hz) for 3; one broadened and one sharp doublet (δP(1) = 23.0 ppm, δP(2) = 64.8; JP(1)-P(2) = 65.0 Hz) for 4. The low-field phosphorus resonances at δ 52.9 and 64.8 ppm were tentatively assigned to the “PPh-µ-(C6H4-ortho)” moiety of 3 and 4, which is connected to boron atom of the cage ligand. The 1H-NMR spectra of 3 and 4, along with the resonances derived from the phenyl protons of the dppp and dppb ligands at δ between ca. 7.8 and 6.8 ppm, as well as the resonances of the cage CH3 protons (δ, 1.14, 0.81 for 3 and δ, 1.09, 0.64 ppm for 4), are indicative of a set of non-equivalent protons of three and four methylene groups attributable to the chelating diphosphine ligands. The latter sets proved to be the only structurally significant and are particularly diagnostic of the presence the ortho-phenylenecycloboronated moiety in 3 and 4. Therefore, we strictly assigned all these resonances by the use of 1H{31P}, 13C{1H} NMR, as well as 2D [13C-1H]-HSQC correlation spectra. Thus, the fully 31P-decoupled 1H{31P} NMR spectra of 3 and 4 each revealed two pairs of apparently simplified 1H multiplet resonances originating from diphosphine CH2 groups substituted by the PPh2 functions. In agreement with this, in the 2D [13C-1H]-HSQC spectrum of 3 (Figure 2, on the left) each pair of these 1H multiplets are connected by cross-peaks to one of the two carbon resonances that appear as doublet of doublets at δ 24.4 and 26.7 ppm with ca. JC,P = 35.3 and JC,P = 7.5 Hz. Similar cross-peaks between each pairs of 1H multiplets, which are susceptible to 31P-decoupling, a broadened doublet at δ 30.8 ppm and a sharp doublet at 25.3 ppm both having ca. same JC,P = 32.0 Hz can be observed in the 2D [13C-1H]-HSQC spectrum of 4 (Figure 2, on the right). The differences in the multiplicity observed for the carbon resonances derived from “Ph2P(1)CH2(CH2)nCH2P(2)Ph-µ-(C6H4-ortho)” units in the 13C{1H} NMR spectra of 3 and 4 can be explained by a shorter carbon chain in the case of diphosphine ligand of 3, which results in additional splitting of resonances derived from CH2 groups on the far-located 31P nucleus.
Figure 2. [1H-13C] HSQC spectrum for 3 (on the left) and 4 (on the right) in CD2Cl2 solution at 25 °C showing the assignment of the proton resonances.
Figure 2. [1H-13C] HSQC spectrum for 3 (on the left) and 4 (on the right) in CD2Cl2 solution at 25 °C showing the assignment of the proton resonances.
Molecules 19 07094 g002

3. Experimental

3.1. General Information

Starting complexes 1 [7] and 2 [9] were prepared according to the published procedures. Chromatography columns (ca. 15 cm in length and 3.5 cm in diameter) packed with silica gel (Acros, Geel, Belgium, 0.035–0.070 mm, 60 A) were used for purification of the prodacts 3 and 4. The 1H, 31P{1H}, 13C{1H}, 11B/11B{1H} and correlation NMR spectra were obtained with a Bruker AMX-400 spectrometer (1H, 400.13 MHz; 31P, 161.98 MHz; 13C, 100.61 MHz; 11B, 128.33 MHz) using TMS as an internal reference and 85% H3PO4 and BF3•Et2O as the external references, respectively. IR spectra were obtained from nujol mull on a Nicolet Magna-750 spectrometer. Microanalyses were performed by Analytical Laboratory of the Institute of Organoelement Compounds of the RAS.

3.2. Preparation of [3-(η2-O2)-3,3,8-[Ph2P(CH2)3PPh-µ-( Molecules 19 07094 i003H8] (3)

A mixture of 1 (113 mg, 0.16 mmol) and an excess of crushed KOH (45 mg, 0.804 mmol) in a 1:1 benzene-methanol mixture (20 mL) was stirred at ambient temperature (~10 min) until all starting material was dissolved and the color of the solution changes from deep red to pale yellow. The flask was opened on air, and the reaction mixture was allowed to stir additionally for 1 h. Solvent was evaporated under reduced pressure and the residue was dissolved in benzene (8–10 mL), filtered and treated by column chromatography on silica gel. The brownish band was eluted from the column using benzene, the solvent was removed, and the resultant solid was re-crystallized from a n-hexane-benzene mixture, affording 50.8 mg (45.2% yield) of pure complex 3 as brown crystals. Anal. Calcd for C31H39B9O2P2Ru: C, 52.89%, H, 5.58%, B, 13.82%. Found: C, 52.82%, H, 5.56%, B, 13.70%. IR (Nujol mull, cm−1): νO-O 941 (w). 1H-NMR (CD2Cl2, 20 °C) δ 7.68 to 6.92 [set of seven m. 19H, Ph2P(1)(CH2)3P(2)Ph-µ-(C6H4-ortho)], 3.01, 2.46 [each m, 1H, Ph2P(1)CH2CH2CH2P(2)Ph-µ-(C6H4-ortho)], 2.90, 2.76 (each m, 1H, Ph2P(1)CH2CH2CH2P(2)Ph-µ-(C6H4-ortho)), 2.38, 1.49 [td and m, 1H, Ph2P(1)CH2CH2CH2P(2)Ph-µ-(C6H4-ortho)], 1.60 (s, H2O in the solvent), 1.14, 0.81 (each s, 3H, CH3-carb). 31P{1H}-NMR (CD2Cl2, 20 °C,) δ 52.9 (d, 1P, JP,P = 75.3 Hz), 18.5 (d, 1P, JP,P = 75.3 Hz). 13C{1H}-NMR (CD2Cl2, 20 °C) δ 139.1, 138.7, 135.5, 128.7 (each d, JC,P = 58.4, 58.7, 54.4, 48.6 Hz, Cipso), 134.8 (br. s), 134.1 (d, JC,P = 6.7 Hz), 134.0 (d, JC,P = 20.0 Hz), 132.2, 130.7, 130.3, 130.2 (each d, JC,P = 2.0, 2.7, 2.2, 2.7 Hz, Cpara), 131.2 (d, JC,P = 7.6 Hz), 128.3 (d, JC,P = 10.8 Hz), 127.8 (d, JC,P = 9.9 Hz), 127.6 (d, JC,P = 9.4 Hz), 126.7 (d, JC,P = 8.6 Hz), 75.6 (br. d-like, Ccarb), 75.2 (br. d-like, Ccarb), 29.7 (s, Me), 26.7 (dd, JC,P = 35.0, 7.0 Hz, CH2P(1)Ph2), 24.4 [dd, JC,P = 35.7, 8.1 Hz, CH2P(2)Ph-µ(C6H4-ortho)], 24.1 (s, Me), 18.3 (s, PCH2CH2CH2P). 11B{1H} NMR (CD2Cl2, 20 °C): 17.2 (s, 1B), −1.8 (3B), −3.0 (1B), −4.0 (1B), −7.9 (3B).

3.3. Preparation of [3-(η2-O2)-3,3,8-[Ph2P(CH2)4PPh-µ-( Molecules 19 07094 i003H8] (4)

The procedure was carried out as for 3: starting complex 1 (45 mg, 0.062 mmol), KOH (17 mg, 0.304 mmol), 1:1 benzene-methanol mixture (10 mL), 10 min + 1 h under air. After isolation and purification via column chromatography on silica gel, the product was then re-crystallized from n-hexane–benzene, affording 20.5 mg (45.8% yield) of pure complex 4 as brownish-yellow crystals. IR (Nujol mull, cm−1): νO-O 938 (w). Anal. Calcd for C32H41B9O2P2Ru•C6H6: C, 57.33%, H, 5.95%. Found: C, 57.62%, H, 6.13%. 1H-NMR (CD2Cl2, 20 °C, J, Hz):, 7.8–6.9 [set of eight overlapped m. 19H, Ph2P(1)(CH2)4P(2)Ph-µ-(C6H4-ortho)], 3.35, 2.37 (m and q-like, 1H, Ph2P(1)CH2CH2CH2CH2P(2)Ph-µ-(C6H4-ortho)), 2.97, 2.63 (q-like and t-like, 1H, Ph2P(1)CH2CH2CH2CH2P(2)Ph-µ-(C6H4-ortho)), 2.05, 1.92, 1.63, 1.27 (each m, 1H, Ph2P(1)CH2CH2CH2CH2P(2)Ph-µ-(C6H4-ortho)), 1.09, 0.64 (each s, 3H, Me). 31P{1H}-NMR (CD2Cl2, 20 °C, J, Hz): 64.7 (d, br, 1P, JP,P = 65.0 Hz), 22.9 (d, 1P, JP,P = 65.0 Hz). 13C{1H}-NMR (CD2Cl2) δ 140.2 (d, J = 54.0 Hz, Cipso), 139.5, 138.9 (overlapping d, JC,P = 51.3, 58.0 Hz, Cipso), 134.3 (d, JC,P = 5.1 Hz), 133.5 (d, JC,P = 20.4 Hz), 131.8, 130.7, 130.2, 130.0 (each d, JC,P = 2.1, 2.1, 2.6, 2.6 Hz, Cpara), 130.9 (d, JC,P = 7.4 Hz), 129.6 (d, JC,P = 41.9 Hz, Cipso), 129.0 (d, JC,P = 4.1 Hz), 128.3 (d, JC,P = 9.5 Hz), 127.7 (d, JC,P = 9.5 Hz), 126.7 (d, JC,P = 8.1 Hz), 76.4–75.8 (m, Ccarb), 30.9 (br. d, JC,P = 32.9 Hz, CH2P(2)Ph-µ-(C6H4-ortho)), 25.4 (d, JC,P = 31.2 Hz, CH2P(1)Ph2), 24.3, 20.0 (each s, PCH2CH2CH2CH2P), 24.2, 20.1 (each s, Me). 11B{1H} NMR (CD2Cl2, 20 °C): 17.5 (s, 1B, B(8)), −1.0 (2B), −2.1 (1B), −3.9 (1B), −6.9 (2B), −7.8 (1B), −10.0 (1B).

3.4. X-ray Structure Determination for Complexes 3 and 4

Single-crystal X-ray diffraction study was carried out with a Bruker SMART 1000 [23] for 3 and Bruker SMART APEX II [24] for 4 diffractometers (graphite monochromated Mo-Kα radiation, λ = 0.71073 Å, ω-scan technique). The SHELXTL software [25] was used for space group and structure determination, refinements, graphics, and structure reporting. The structures were solved by direct methods and refined by the full-matrix least-squares technique against F2 with the anisotropic thermal parameters for all non-hydrogen atoms. The hydrogen atoms of carborane moieties were located from difference Fourier maps, the rest hydrogen atoms were placed in the geometrically calculated positions. All hydrogen atoms were included in the structure factors calculations in the riding motion approximation. In the crystal structure of 4, the central atom C (16B) of dppb ligand in the one of two independent molecules is disordered over two positions with 0.8/0.2 occupancies. In both structures 3 and 4, the CH2Cl2 solvent molecules are presented. The principal experimental and crystallographic parameters are listed in Table 2. CCDC 992977 (for 3) and 992978 (for 4) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44 1223 336033; E-mail: [email protected]).
Table 2. Crystal data, data collection and structure refinement parameters for 3 and 4.
Table 2. Crystal data, data collection and structure refinement parameters for 3 and 4.
Compound34
Molecular formulaC31H39B9O2P2Ru•2(CH2Cl2)C32H41B9O2P2Ru•0.25(CH2Cl2)
Formula weight873.77739.18
Dimension, mm30.20 × 0.15 × 0.100.28 × 0.14 × 0.07
Crystal systemtriclinictriclinic
Temperature, K120100
Space groupP?1P?1
a, Å9.791(2)10.361(2)
b, Å10.063(4)16.496(4)
c, Å20.831(1)21.721(5)
α, deg.82.764(1)76.958(4)
β, deg.84.678(1)81.790(4)
γ, deg.62.696(1)79.194(4)
V, Å31894.6(1)3533(1)
Z24
ρcalc, g·cm−31.5321.390
2 θmax, deg.6054
Linear absorption ( µ), cm−18.137.95
No. unique refl. ( Rint)10975 (0.0306)15322 (0.1018)
No. observed refl. ( I > 2σ(I))82548960
No. parameters470877
R1 (on F for observed refl.) a0.04560.0534
wR2 (on F2 for all refl.) b0.10170.1197
GOOF0.9800.907
Note: a R1 = Σ||Fo − |Fc||/Σ|Fo|; b wR2 = {Σ[w(Fo2Fc2)2]/Σw(Fo2)2}1/2.

4. Conclusions

The complexes described in this paper are novel oxygen-containing ortho-phenylenecycloboronated closo-ruthenacarboranes of the general formula [3-(η2-O2)-3,3,8-[Ph2P(CH2)nPPh-µ-( Molecules 19 07094 i003H8] (3, n = 3) and (4, n = 4). They were prepared by a facile and potentially useful method starting from paramagnetic species [3-Cl-3,3,8-{Ph2P(CH2)nPPh-µ-( Molecules 19 07094 i002H8] (n = 3 and 4), which formally undergo a mild base-mediated chlorine-oxygen displacement reaction. Based on these and the literature data, it is envisaged that suitably constituted paramagnetic ruthenacarborane complexes as the potential precursors might open up extensive applications in cluster ruthenacarborane chemistry as well as in the homogeneous metallacarborane catalysis.

Acknowledgments

This work has been supported by Grants Nos 12-0300102 and 14-0331618 from Russian Foundation for Basic Research. We thank M.G. Esernizkaya for recording IR spectra of 3 and 4.

Author Contributions

Alexander Y. Kostukovich and Dmitrii I. D’yachihin performed research and analyzed the data, Fedor M. Dolgushin and Alexander F. Smol’yakov performed X-Ray diffraction studies, Ivan A. Godovikov provided NMR recording, Igor T. Chizhevsky supervised the work and engaged in writing manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Crook, J.E.; Greenwood, N.N.; Kennedy, J.D.; McDonald, W.S. A novel exo-bicyclic closo-six-vertex dimetalla-hexaborane: [1,1,2-(CO)3–1-(PPh3)-2,2-(Ph2P Molecules 19 07094 i004H2–3,5-)]•CH2Cl2. J. Chem. Soc. Chem. Commun. 1982. [Google Scholar] [CrossRef]
  2. Elrington, M.; Greenwood, N.N.; Kennedy, J.D.; Thornton-Pett, M. Preparation and nuclear magnetic resonance properties of eleven-vertex closo-type osmaundecaboranes and the X-ray crystal structure of the ortho-cycloboronated compound [2,5-(OEt)2-1-(PPh3)-1-(o-Ph2P Molecules 19 07094 i005]. J. Chem. Soc. Dalton Trans. 1986. [Google Scholar] [CrossRef]
  3. Bould, J.; Brint, P.; Kennedy, J.D.; Thornton-Pett, M. Metallaborane reaction chemistry. Part 1. Two interesting closed cluster compounds from the reaction of acetylene with an open nido-6-iridadecaborane. J. Chem. Soc. Dalton Trans. 1993. [Google Scholar] [CrossRef]
  4. Crook, J.E.; Greenwood, N.N.; Kennedy, J.D.; McDonald, W.S. A novel oxidative insertion of a metal centre into a degraded closo-borane cluster; crystal and molecular structure of the ten-vertex iso-nido cluster [{IrC(OH) Molecules 19 07094 i006H4PPh2)(PPh3)]. J. Chem. Soc. Chem. Commun. 1981. [Google Scholar] [CrossRef]
  5. Pisareva, I.V.; Konoplev, V.E.; Balagurova, E.V.; Yanovsky, A.I.; Dolgushin, F.M.; Cheredilin, D.N.; Chizhevsky, I.T.; Franken, A.; Carr, M.J.; Thornton-Pett, M.; et al. Metal-assisted polyhedral contraction reactions of monocarbon and dicarbon carborane species involving nido-{CB10}, nido-{CB9}, nido-C2B8} and exo-nido-C2B9Os} systems. In Boron Chemistry at the Beginning of the 21st Century; Bubnov, Y., Ed.; Editorial URSS: Moscow, Russian, 2003; pp. 271–278. [Google Scholar]
  6. Cheredilin, D.N.; Kadyrov, R.; Dolgushin, F.M.; Balagurova, E.V.; Godovikov, I.A.; Solodovnikov, S.P.; Chizhevsky, I.T. Chiral paramagnetic closo-ruthenacarboranes via phosphine-diphosphine displacement reaction of “three-bridge” exo-nido-ruthenacarboranes: Molecular Structure of (−)-[closo-3-Cl-3,3-{(Ph2PCHCH3)2CH2}-3,1,2-RuC2B9H11] and its ortho-cycloboronated derivative. Inorg. Chem. Commun. 2005, 8, 614–618. [Google Scholar]
  7. Kolyakina, E.V.; Grishin, I.D.; Cheredilin, D.N.; Dolgushin, F.M.; Chizhevsky, I.T.; Grishin, D.F. Ruthenium carborane complexes in the controlled radical polymerization of methyl methacrylate. Russ. Chem. Bull. 2006, 55, 89–93. [Google Scholar] [CrossRef]
  8. Grishin, I.D.; D’yachihin, D.I.; Piskunov, A.V.; Dolgushin, F.M.; Smol’yakov, A.F.; Il’in, M.M.; Davankov, V.A.; Chizhevsky, I.T.; Grishin, D.F. Carborane complexes of ruthenium(III): Studies on thermal reaction chemistry and the catalyst design for atom transfer radical polymerization of methyl methacrylate. Inorg. Chem. 2011, 50, 7574–7585. [Google Scholar] [CrossRef]
  9. Grishin, I.D.; Turmina, E.S.; D’yachihin, D.I.; Vinogradov, D.S.; Piskunov, A.V.; Smol’yakov, A.F.; Dolgushin, F.M.; Chizhevsky, I.T.; Grishin, D.F. Efficient catalytic systems based on paramagnetic closo-ruthenacarboranes for the controlled synthesis of polymers. Russ. Chem. Bull. 2011, 60, 2375–2383. [Google Scholar] [CrossRef]
  10. Grishin, I.D.; Turmina, E.S.; D’yachihin, D.I.; Chizhevsky, I.T.; Grishin, D.F. Carborane complexes of ruthenium with long-chain diphosphine ligands as effective catalysts of controlled radical polymerization. Polym. Sci. B 2014, 56, 1–10. [Google Scholar]
  11. Grishin, I.D.; D’yachihin, D.I.; Turmina, E.S.; Dolgushin, F.M.; Smol’yakov, A.F.; Piskunov, A.V.; Chizhevsky, I.T.; Grishin, D.F. Mononuclear closo-ruthenacarborane complexes containing a rare eight-membered metal-diphosphine ring. J. Organomet. Chem. 2012, 721–722, 113–118. [Google Scholar] [CrossRef]
  12. We note that a series of some other closo-bis(phosphine)hydridoruthenacarboranes with monoanionic charge-compensated cage ligands [13] were also found to display very high activity as Kharasch catalysts [14], where the importance of Ru (II) → Ru (III) transformation in this catalytic process is stressed.
  13. Tatusaus, O.; Núñez, R.; Viñas, C.; Teixidor, F.; Mata, I.; Molins, E. Synthesis, characterization, and dinamic studies of 12-vertex η5-ruthenium(II) closo-Phosphine complexes with monoanionic [10-L-nido-7-R-7,8-C2B9H9]-ligands. Inorg. Chem. 2004, 43, 6067–6074. [Google Scholar] [CrossRef]
  14. Tatusaus, O.; Viñas, C.; Núñez, R.; Teixidor, F.; Demonceau, A.; Delfosse, S.; Noels, A.F.; Mata, I.; Molins, E. The modulation possibilities of dicarbollide clusters: Optimizing the Kharasch catalysts. J. Am. Chem. Soc. 2003, 125, 11830–11831. [Google Scholar] [CrossRef]
  15. 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]
  16. Jones, C.J.; Francis, J.N.; Hawthorne, M.F. The chemistry of an 11-atom polyhedral metallocarborane prepared by polyhedral contraction. J. Chem. Soc. Chem. Commun. 1972. [Google Scholar] [CrossRef]
  17. Dustin, D.F.; Hawthorne, M.F. Synthesis of nine-vertex, monocarbon metallocarboranes by polyhedral contraction. Inorg. Chem. 1973, 12, 1380–1383. [Google Scholar] [CrossRef]
  18. Hanusa, T.P.; Huffman, J.C.; Curtis, T.L.; Todd, L.J. Synthesis of (π-arene)metallacarbaboranes containing ruthenium or osmium and a (π-cyclohexadienyl)cobaltacarbaborane. Crystal structure of 2,5,6-(η-C6H6)RuC2B7H11. Inorg. Chem. 1985, 24, 787–792. [Google Scholar] [CrossRef]
  19. Kirchner, K.; Mauthner, K.; Mereiter, K.; Schmid, R. Irreversible binding of dioxygen and other gases to a half-sandwich ruthenium(II) complex: X-ray structure of [Ru(η2-O2)(η5-C5Me5)(Ph2PCH2CH2PPh2)]PF6. J. Chem. Soc. Chem. Commun. 1993. [Google Scholar] [CrossRef]
  20. Jia, G.; Ng, W.S.; Chu, H.S.; Wong, W.-T.; Yu, N.-T.; Williams, I.D. Dioxygen Complexes from the Reactions of [Cp*RuH2(PP)]+ (PP = dppm, dppe) with Air. Organometallics 1999, 18, 3597–3602. [Google Scholar] [CrossRef]
  21. Teixidor, F.; Ayllón, J.A.; Viñas, C.; Sillanpää, R.; Kivekäs, R.; Casabó, J. Iridium coordination to exo-dithio-7,8-dicarba-nido-undecaborate derivatives. Inorg. Chem. 1994, 33, 4815–4818. [Google Scholar] [CrossRef]
  22. Yanovsky, A.I. Metallocarbaboranes. Structural studies in recent years. Russ. Chem. Rev. 1985, 54, 515–531. [Google Scholar] [CrossRef]
  23. SMART v5.051 and SAINT v 5.00, Area Detector Control and Integration Software, Bruker AXS Inc.: Madison, WI, USA, 1998.
  24. APEX II Software Package, Bruker AXS Inc.: Madison, WI, USA, 2005.
  25. Sheldrick, G.M. A short history of SHELX. Acta Cryst. 2008, A64, 112–122. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds 14 are available from the authors.

Share and Cite

MDPI and ACS Style

Kostukovich, A.Y.; D'yachihin, D.I.; Dolgushin, F.M.; Smol'yakov, A.F.; Godovikov, I.A.; Chizhevsky, I.T. An Unusual Conversion of Paramagnetic [3-Cl-3,3,8-{Ph2P(CH2)nPPh-µ-(C6H4-ortho)}-1,2-(CH3)2-closo-3,1,2-RuIIIC2B9H8] (n = 3 and 4) to Form the First 18-Electron P-Phenylene ortho-Cycloboronated closo-Ruthenacarboranes with a Dioxygen Ligand. Molecules 2014, 19, 7094-7103. https://doi.org/10.3390/molecules19067094

AMA Style

Kostukovich AY, D'yachihin DI, Dolgushin FM, Smol'yakov AF, Godovikov IA, Chizhevsky IT. An Unusual Conversion of Paramagnetic [3-Cl-3,3,8-{Ph2P(CH2)nPPh-µ-(C6H4-ortho)}-1,2-(CH3)2-closo-3,1,2-RuIIIC2B9H8] (n = 3 and 4) to Form the First 18-Electron P-Phenylene ortho-Cycloboronated closo-Ruthenacarboranes with a Dioxygen Ligand. Molecules. 2014; 19(6):7094-7103. https://doi.org/10.3390/molecules19067094

Chicago/Turabian Style

Kostukovich, Alexander Y., Dmitrii I. D'yachihin, Fedor M. Dolgushin, Alexander F. Smol'yakov, Ivan A. Godovikov, and Igor T. Chizhevsky. 2014. "An Unusual Conversion of Paramagnetic [3-Cl-3,3,8-{Ph2P(CH2)nPPh-µ-(C6H4-ortho)}-1,2-(CH3)2-closo-3,1,2-RuIIIC2B9H8] (n = 3 and 4) to Form the First 18-Electron P-Phenylene ortho-Cycloboronated closo-Ruthenacarboranes with a Dioxygen Ligand" Molecules 19, no. 6: 7094-7103. https://doi.org/10.3390/molecules19067094

Article Metrics

Back to TopTop