Next Article in Journal
Molecularly Engineered Lithium–Chromium Alkoxide for Selective Synthesis of LiCrO2 and Li2CrO4 Nanomaterials
Next Article in Special Issue
Mono- and Hexanuclear Zinc Halide Complexes with Soft Thiopyridazine Based Scorpionate Ligands
Previous Article in Journal
Modelling the (Essential) Role of Proton Transport by Electrolyte Bases for Electrochemical Water Oxidation at Near-Neutral pH
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Trithia-Borinane Complexes Stabilized in Diruthenium Core: [(Cp*Ru)21-S)(η1-CS){(CH2)2S3BR}] (R = H or SMe)

Department of Chemistry, Indian Institute of Technology Madras, Chennai 600036, TN, India
*
Author to whom correspondence should be addressed.
Submission received: 12 December 2018 / Revised: 5 February 2019 / Accepted: 7 February 2019 / Published: 13 February 2019
(This article belongs to the Special Issue Metal Complexes Containing Boron Based Ligands)

Abstract

:
The thermolysis of arachno-1 [(Cp*Ru)2(B3H8)(CS2H)] in the presence of tellurium powder yielded a series of ruthenium trithia-borinane complexes: [(Cp*Ru)21-S)(η1-CS){(CH2)2S3BH}] 2, [(Cp*Ru)21-S)(η1-CS){(CH2)2S3B(SMe)}] 3, and [(Cp*Ru)21-S)(η1-CS){(CH2)2S3BH}] 4. Compounds 24 were considered as ruthenium trithia-borinane complexes, where the central six-membered ring {C2BS3} adopted a boat conformation. Compounds 24 were similar to our recently reported ruthenium diborinane complex [(Cp*Ru){(η2-SCHS)CH2S2(BH2)2}]. Unlike diborinane, where the central six-membered ring {CB2S3} adopted a chair conformation, compounds 24 adopted a boat conformation. In an attempt to convert arachno-1 into a closo or nido cluster, we pyrolyzed it in toluene. Interestingly, the reaction led to the isolation of a capped butterfly cluster, [(Cp*Ru)2(B3H5)(CS2H2)] 5. All the compounds were characterized by 1H, 11B{1H}, and 13C{1H} NMR spectroscopy and mass spectrometry. The molecular structures of complexes 2, 3, and 5 were also determined by single-crystal X-ray diffraction analysis.

Graphical Abstract

1. Introduction

The mutually synergistic interactions between metals and organic ligands often generate compounds of fundamental and practical importance [1,2,3,4,5,6]. The structure and reactivity of metallaboranes, which features compounds with an M–B bond, is greatly influenced by transition metals as well as organic ligands [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25]. Previous studies have been carried out to understand the ways in which metal and borane fragments can interact to generate novel geometries [1,2,3,4,16,17,18,19,20,21,22,23,24,25]. However, there is still little understanding of how a transition metal can be used to vary the chemistry of metallaborane compounds. In this regard, our group was actively involved in the synthesis of various electron-precise transition metal–boron complexes such as σ-borane [26,27,28,29,30,31], boryl [32,33], triply-bridged trimetallic borylene [34,35,36,37,38], diborane [39], B-agostic [26,27,40,41,42], and metallaboratrane [26,27,43,44] complexes using of different synthetic precursors. An important aspect is the incorporation of transition metals into the chemistry of p-block elements other than carbon [45,46,47]. The literature contains numerous examples for boron, but other elements illustrate the possibilities as well [48,49]. The chemistry of transition-metal complexes with main group elements, particularly with chalcogen ligands, are of substantial importance. The homo- and heterometallic sulfido complexes with a wide range of substrates are well-documented in the literature [50,51,52,53]. In contrast, thioborates are not regularly seen in the coordination sphere of transition metals, mostly due to the lack of synthetic routes. It is interesting to see how a change of metal or ligand plays an important role in determining the nature of the molecules (Chart 1).
Several research groups have explored this idea, which has led to the isolation of unique molecules with interesting bonding interactions [1,54,55,56,57,58,59,60,61,62,63,64,65]. Here, we have tried to provide a quick overview of several such examples reported by us and others [26,54,55,56,57,58,59,60,61,62,63,64,65]. Hartwig in 1996 reported the first example of a σ-borane metal complex, I, from the reaction of catecholborane and dimethyl titanocene [1]. Following this, several research groups were successful in isolating σ-borane/borate complexes [54,55,56]. Weller and colleagues synthesized a novel bis(σ-amine–borane) complex of rhodium through the displacement of a labile fluoroarene ligand from [Rh(η6-C6H5F){P(C5H9)22-C5H7)}][BArF4] [54]. Inspired by this, our group recently reported a σ-borane complex of ruthenium from the reaction of ruthenium bis(σ)borate and [Mn2(CO)10] [26,27]. The first metalladiborane [(η5-C5H8)Fe(CO)22-B2H5)], II, was structurally characterized by Shore in 1989 [57,58]. We recently reported a ruthenium diborane, a derivative of diborane(6) from the reaction of [(Cp*Ru)2B3H9] (Cp* = η5-C5Me5) and 2-mercaptobenzothiazole [26]. Sabo-Etienne and colleagues have recently shown the formation of a ruthenium agostic complex [RuH2{η2-H-B(NiPr2)-CH2PPh2}(PCy3)2], VII, by treating phosphinomethyl(amino)borane [Ph2PCH2BHNiPr2] and [RuH2(η2-H2)2(PCy3)2] [59]. The reaction of Na[(H2B)mp2] (mp = 2-mercaptopyridyl) and [Re2CO10] enabled us to isolate an agostic complex of rhenium, [Re(CO)3(μ-H)BH(C5H4NS)2], X [27]. Hill and colleagues established how scorpionate ligands can be utilized for the formation of complexes that have a direct metal boron bond through the isolation of the first metallaboratrane, [M(CO)(PPh3){B(mt)3}](M→B) (mt = methimazolyl, M = Ru and Os) in 1999 [60]. Following this, Bourissou and Parkin synthesized a RhI metallaboratrane [61], XII, and a ferraboratrane [{k4-B(mimtBu)3}Fe(CO)2] (mimtBu = 2-mercapto-1-tert-butylimidazolyl) [62], XIII, respectively. We successfully isolated a ruthenaboratrane by using [(η6-p-cymene)RuCl2]2 as a precursor XIV [43], whereas a rhoda/irida boratrane, [Cp*M(BHL2)], (L = C5H4NS, M = Rh or Ir) [43], XV, could be synthesized from the reaction of [Cp*MCl2]2 with Na[H2B(mp)2]. Marder and colleagues synthesized metal-bridged-boryl complexes by using catecholborane [63]. In 2005, Braunschweig reported a heterometallic Fe–Pd bridged-boryl complex from the reaction of [Cp*Fe(CO)2BCl2] and [Pd(Cy3)2] [64]. Later, our group successfully synthesized a homometallic ruthenium bridged-boryl complex from the reaction of HBcat (catecholborane, cat = 1,2-O2C6H4) and [{Ru(CO)}2B2H6] [32]. Following this, we recently reported a bis(bridging-boryl) complex, [{Cp*Ru(µ,η2-HBS2CH2)}2], from the thermolysis of [Cp*Ru(µ-H)2BH(S-CH=S)] with chalcogen powder [33]. Fehlner and colleagues reported a homometallic bridging borylene complex XVIII [65] from the reaction of [CpCo(PPh3)2] and BH3·THF. Our group was successful in synthesizing heterometallic triply bridged borylene complexes [(Cp*Co)23-BH)(μ-CO){M(CO)5}] (M = W, Mo, Cr) from the reaction of [{Cp*CoCl}2] and LiBH4·THF with [M(CO)3(MeCN)3] [34,35,36,37,38].
Ligands such as COS, CS2, and CO2 interact with transition metal complexes, showing a wide range of chemical transformations, such as insertion, dimerization, disproportionation, coupling, and catalytic reactions [66,67,68]. On the basis of the general concern of the electron donating/accepting properties of CS2 and CO2, various binding modes with one or more metal atoms have been recognized [69]. However, reactivities of these ligands towards polyhedral metallaborane clusters have been sparsely explored [70,71,72,73,74]. In this context, Fehlner and colleagues described the reactivity of CS2 with an unsaturated chromaborane cluster [(Cp*Cr)2B4H8], which underwent metal-assisted hydroboration and successively converted to a methanedithiolato ligand [71]. Following this, our group reported the reaction of CS2 with nido-[(Cp*Ru)2(μ-H)2B3H7], which subsequently transformed into [(Cp*Ru)2(B3H8)(CS2H)], 1, containing a dithioformato ligand (CHS2) [69]. Recently, we reported for the first time a ruthenium trithia-diborinane complex, 1-thioformyl-2,6-tetrahydro-1,3,5-trithia-2,6-diborinane [(Cp*Ru){(η2-SCHS)CH2S2(BH2)2}], from the reaction of [{Cp*RuCl(µ-Cl)}2] and Na[BH3(SCHS)] [33]. Encouraged by these results, we became interested in exploring the reactivity of 1 under different reaction conditions, especially with heavier chalcogen ligands. Thus, we performed the reaction of 1 in the presence of chalcogen powder. As expected, the reaction enabled us to isolate some interesting ruthenium trithia-borinane complexes.

2. Results and Discussion

Synthesis of Ruthenium Borinane Complexes, 24

As shown in Scheme 1, the pyrolysis of 1 in the presence of tellurium powder in toluene yielded compounds 24 along with compounds [{Cp*Ru(µ,η3-SCHS)}2] and [Cp*Ru(μ-H)2BH(SCHS)] [33]. The 11B{1H} NMR spectra at room temperature display single resonance at δ = −4.1, 7.4, and 4.9 ppm for compounds 2, 3, and 4, respectively, indicating the presence of a single boron atom. While the 1H NMR spectrum of compounds 2 and 4 shows the presence of a terminal B–H proton at δ = 3.75 and 2.58 ppm, respectively, compound 3 does not show any indication of a B–H terminal. Instead, it shows a resonance at δ = 2.06 ppm, indicating the presence of a (SCH3) unit. Apart from that, both 2 and 3 display resonances in the region δ = 3.96–1.69 ppm, which may be attributed to the presence of methylene protons. Both compounds display signals for two sets of Cp* protons around 1.79 and 1.72 ppm in a 1:1 ratio. The presence of the Cp* ligands, methylene, and SCH3 units are also supported by 13C{1H} NMR spectroscopy. Apart from that, the 13C{1H} NMR spectra also show a resonance at δ = 288.6 and 285.8 ppm, indicating the presence of a C=S group in the molecules of 2 and 3 respectively. Furthermore, the mass spectra show molecular ion peaks (ESI+) at m/z = 686.9603, 732.9479, and 686.9604 for compounds 2, 3, and 4 respectively. Although we isolated the majority of Te powder after workup, we are not in a position to comment on the exact role of chalcogen powder, in particular Te powder, in the formation of complexes 24 from 1.
The single-crystal X-ray diffraction study disclosed the core geometry (C2S3B ring) of compounds 2 and 3 to be very similar to each other (Figure 1a,b). The only difference between the two is the position of the boron atom in the central six-membered ring {C2S3B}. Compounds 2 and 3 can be called as 1,3,5-trithia-4-borinane and 1,3,5-trithia-2-borinane complexes of ruthenium, respectively, which is similar to our recently reported diborinane [(Cp*Ru){(η2-SCHS)CH2S2(BH2)2}] [33]. Unlike diborinane, compounds 2 and 3 have only one boron atom in the six-membered ring {C2S3B} and are the monoborane derivatives of [(Cp*Ru){(η2-SCHS)CH2S2(BH2)2}]. While the central six-membered ring adopts a chair conformation in diborinane [33], 2 and 3 adopt a boat conformation. A significant difference between 2 and 3 is the presence of the {SMe} moiety instead of a terminal hydrogen attached to the boron atom in compound 3. The B–S bond length (av. 1.921 Å) in 2 and 3 is within the B–S single bond distance and is in accord with the ruthenium diborinane complex [33]. One of the interesting features observed in these molecules is the presence of the thioformyl unit bonded to the ruthenium atoms. While the diborinane has only one ruthenium atom, compounds 2 and 3 has two ruthenium atoms bridged by one thiocarbonyl unit on one side and B–S on the other side. The C–S distance in the thiocarbonyl unit (1.612(15) Å in 2 and 1.617(7) Å in 3) is found to be shorter than that of 1. The Ru1–Ru2 distances of 2.759(6) Å in 2 and 2.759(6) Å in 3 are significantly shorter when compared to 1, but are well within the reported Ru–Ru single bond distance [69]. The ruthenium atoms are connected to two sulfur atoms S2 and S4 present in the (C2S3B) ring and the bridging sulfur is connected to the ring boron atom B1. Although we failed to crystallize compound 4, it was characterized in comparison to its spectroscopic data with 2 and 3. Based on the spectroscopic data, compound 4 is expected to have a structure similar to that of compound 3 where instead of the SMe group, a terminal H is attached to the B atom (Scheme 1).
The six-membered ring containing a {C2BS3} moiety adopts a boat conformation, similar to the reported diborinanes, such as bis(cAAC)-stabilized 3,6-dicyano-1,2,4,5-tetrasulfa-3,6-diborinane reported by Braunschweig et al. where the central {B2S4} ring displayed a boat conformation and was the first example of a structurally and NMR-spectroscopically characterized {B2S4}-heterocycle [75]. Meller et al. reported the synthesis and characterization of a diborinane-tungsten adduct, [(BMe)2(NH){N(SiMe3)}2(S){W(CO)5}] [76]. In contrast, the structurally characterized dioxaborinane, [CN(C6H5)(BO2C3H5)(C6H4)(C4H9)], adopted the half-chair conformation [77]. Recently, our group reported for the first time a trithia-diborinane stabilized ruthenium complex, [(Cp*Ru){(η2-SCHS)CH2S2(BH2)2}] [33]. Although some examples of trithia-diborinane compounds have been reported, there are no examples of metal complexes of such trithia-diborinane species except the one reported by us [33]. Compounds 24 are the monoborinane derivatives, and are a novel entry to the class of transition metal borinane complexes. The few structurally characterized borinane and diborinane derivatives are listed in Table 1.
In order to check whether arachno-[(Cp*Ru)2(B3H8)(CS2H)], 1, can be converted to a nido or closo geometry with the release of hydrogen, we pyrolyzed 1. Interestingly, the reaction led to the formation of 5 having a capped butterfly geometry, instead of a nido or closo geometry (Scheme 2). The mass spectrometry of the new compound gives a molecular ion peak at m/z = 613.0588 that corresponds to C21H37Ru2B3S2Na. The room-temperature 11B{1H} NMR spectrum of 5 rationalizes the presence of two boron environments, which appear at δ = 43.6 and −24.1 ppm. Besides the BH terminal protons, one B–H–B and one Ru–H–B proton is observed in the 1H NMR spectrum. Furthermore, the 1H NMR spectrum implies the presence of two equivalent Cp* ligands in 5.
The identity of 5 is confirmed by its solid-state X-ray crystal analysis. The asymmetric unit of 5 contains two independent molecules and the structural data presented here are from one of the units (Figure 2). In one of the units, the B5–B4–B3–S4–C43–S3 moiety is disordered over two positions with occupancy factors 0.602 and 0.398. As shown in Figure 2, the molecular structure of 5 can be viewed as a capped butterfly cluster, where one of the triangular faces (Ru1–B2–Ru2) is capped by a BH fragment (B1 in Figure 2). The Ru1–Ru2 distance in 5 is shorter than that observed in 1 by 0.258 Å. While the Ru–B distances in both 1 and 5 is comparable, the B–B distances show considerable variation. It is worth noting that the B1–B2 bond distance of 1.679(11) Å in 5 is shorter than the normal B–B single bond, but it is comparable to that of a manganese hexahydridodiborate complex [{(OC)4Mn}(η6-B2H6){Mn(CO)3}2(µ-H)] [39]. The interatomic separation between B3 and S1 (3.029 Å) is significantly longer for the formation of a direct B–S bond, and is bridged via the {S-CH2} unit. With seven-skeletal-electron-pairs (sep), compound 5 satisfies the electron count for a BH capped arachno-butterfly structure. By the fused polyhedral model of Mingos [79,80,81,82], 5 should have 44 electrons [Ru2B2 (butterfly); 42 + Ru2B2 (tetrahedron); 40 – Ru2B (face); 38], which is also supported by the cve count of 44 electrons [2 (Cp*Ru) × 13 + 1 (μ2-S) × 1 + 1 × (μ3-S) × 3 + 3 (BH) × 4 + 2 (H) × 1]. Compound 5 thus obeys the Wade–Mingos rule for an arachno system [79,80,81,82].

3. Materials and Methods

3.1. General Procedures and Instrumentation

All manipulations were conducted under an Ar/N2 atmosphere using standard Schlenk techniques or glove box techniques. The solvents were distilled prior to use under argon. Compound arachno-1 was prepared according to the literature method [69], while other chemicals were obtained commercially and used as received. The external reference [Bu4N][B3H8] for the 11B NMR was synthesized with the literature method [83]. Preparative thin layer chromatography was performed with Merck 105554 silica-gel TLC plates (Merck, Darmstadt, Germany). The NMR spectra were recorded on a 400 or 500 MHz Bruker FT-NMR spectrometer (Bruker, Billerica, MA, USA). Residual solvent protons were used as reference (δ, ppm CDCl3, 7.26), while a sealed tube containing [Bu4N(B3H8)] in [d6]-benzene (δB, ppm, −30.07) was used as an external reference for the 11B NMR. The FT-IR spectrum was recorded using a Jasco FT/IR-4100 spectrometer (JASCO, Easton, MD, USA). The HRMS (ESI) spectra were obtained using a Bruker Micro TOF-II instrument (Bruker, Billerica, MA, USA). Note that all the reported compounds were isolated by the preparative thin layer chromatographic technique (TLC), using silica-gel-coated aluminum TLC plates. The impure reaction mixture was slowly loaded on the TLC and eluted by using the hexane/CH2Cl2 mixture in inert atmosphere. Elution with the particular solvent mixture allowed us to separate the compounds in pure form.

3.2. Synthesis

3.2.1. Synthesis of Compounds 2, 3, and 4

In a flame-dried Schlenk tube, compound 1 (0.1 g, 0.169 mmol) was suspended in toluene (20 mL), and Te powder (0.58 g, 0.97 mmol) was added. The reaction mixture was stirred for 24 h at 80 °C. The solvent was evaporated in vacuum, then the residue was extracted into hexane/CH2Cl2 (60:40 v/v) and passed through Celite. After the removal of the solvent from the filtrate, the residue was subjected to chromatographic workup using silica-gel TLC plates. Elution with hexane/CH2Cl2 (60:40 v/v) yielded pink solid 2 (0.012 g, 10%), pink solid 3 (0.009 g, 7%), and pink solid 4 (0.008 g, 7%) along with the compounds [{Cp*Ru(µ,η3-SCHS)}2] (0.002 g, 2%) and [Cp*Ru(μ-H)2BH(SCHS)] (0.003 g, 4%).
2: HR-MS (ESI+) calcd. for C23H36S5BRu2+ [M + H]+ m/z 686.9601, found 686.9603; 11B{1H} NMR (160 MHz, CDCl3, 22 °C): δ = −4.1 ppm (br, 1B); 1H NMR (500 MHz, CDCl3, 22 °C): δ = 3.81, 2.94, 2.01, 1.70 (d, 4H, CH2S2), 3.75 (br, 1H, BHt,), 1.74, 1.72 (s, 30H, 2 × Cp*); 13C{1H} NMR (125 MHz, CDCl3, 22 °C): δ = 288.6 (s, CS), 96.5, 96.4 (s, C5Me5), 28.7, 11.8 (s, CH2S2), 9.8, 9.4 ppm (s, C5Me5); IR (CH2Cl2): ν ˜ = 2494 (BHt), 1089 cm−1 (µ-CS).
3: HR-MS (ESI+) calcd for C24H38BS6Ru2+ [M + H]+ m/z 732.9478, found 732.9479; 11B{1H} NMR (160 MHz, CDCl3, 22 °C): δ = 7.4 ppm (br, 1B); 1H NMR (500 MHz, CDCl3, 22 °C): δ = 3.97, 3.17, 2.19, 1.82 (d, 4H, CH2S2), 2.05 (s, 3H, SCH3), 1.79, 1.73 (s, 30H, 2 × Cp*); 13C{1H} NMR (125 MHz, CDCl3, 22 °C): δ = 285.8 (s, CS), 97.3, 96.5 (s, C5Me5), 35.3, 17.1 (s, CH2S2), 12.7 (s, SCH3), 10.1, 9.4 ppm (s, C5Me5); IR (CH2Cl2): ν ˜ = 1085 cm−1 (µ-CS).
4: HR-MS (ESI+) calcd for C23H36BS5Ru2+ [M + H]+ m/z 686.9601, found 686.9604; 11B{1H} NMR (160 MHz, CDCl3, 22 °C): δ = 4.9 ppm (br, 1B); 1H NMR (500 MHz, CDCl3, 22 °C): δ = 3.93, 3.17, 2.20, 1.76 (d, 4H, CH2S2), 2.58 (br, 1H, BHt), 1.80, 1.73 (s, 30H, 2 × Cp*); 13C{1H} NMR (125 MHz, CDCl3, 22 °C): δ = 97.3, 96.5 (s, C5Me5), 35.3, 17.1 (s, CH2S2), 10.1, 9.4 ppm (s, C5Me5); IR (CH2Cl2): ν ˜ = 2383 cm−1 (BHt), 1081 cm−1 (µ-CS).

3.2.2. Synthesis of Compound 5

In a flame-dried Schlenk tube, compound 1 (0.1 g, 0.169 mmol) was suspended in toluene (20 mL), and was stirred at 80 °C for 18 h. The solvent was evaporated in vacuum, and the residue was extracted into hexane/CH2Cl2 (70:30 v/v) and passed through Celite. After the removal of the solvent from the filtrate, the residue was subjected to chromatographic workup using silica-gel TLC plates. Elution with hexane/CH2Cl2 (70:30 v/v) yielded orange 5 (0.030 g, 30%).
5: HR-MS (ESI+) calcd for C21H37B3NaS2Ru2+ [M + Na]+ m/z 613.0601, found 613.0588; 11B{1H} NMR (160 MHz, CDCl3, 22 °C): δ = 43.6, −24.1 ppm (br, 2B); 1H NMR (500 MHz, CDCl3, 22 °C): δ = 5.09 (br, 3H, BHt) 3.89, 2.94 (d, 2H, CH2S2), 1.86, 1.81 (s, 30H, 2 × Cp*), −2.08 (br, 1H, B–H–B), −13.41 (br, 1H, Ru–H–B); 13C{1H} NMR (125 MHz, CDCl3, 22 °C): δ = 95.8, 92.2 (s, C5Me5), 41.1 (s, CH2S2), 11.7, 11.1 ppm (s, C5Me5); IR (CH2Cl2): ν ˜ = 2450 (BHt), 2046 (Ru–H–B).

3.3. X-ray Crystallography

The crystal data for compounds 2, 3, and 5 were collected and integrated using a Bruker APEX II CCD diffractometer (Bruker, Billerica, MA, USA), with graphite monochromated Mo-Kα (λ = 0.71073 Å) radiation at 296 K (2 and 3) and 293 K (5). The structures were solved by heavy atom methods using SHELXS-97 [84] and refined using SHELXL-2013 for compound 2 and SHELXL-2014 [85] for compound 3. The structure of compound 5 was solved by heavy atom method using SIR-92 [86] and SHELXL-2014. The crystallographic data were deposited at the Cambridge Crystallographic Data Centre as Supplementary Materials no. CCDC-1856640 (2), CCDC-1828322 (3), and CCDC-1407806 (5). These data can be obtained free-of-charge from the Cambridge Crystallographic Data Center via www.ccdc.cam.ac.uk/data_request/cif.
Crystal data for compound (2): C23H35BRu2S5, Mr = 684.76, monoclinic, space group C2/c, a = 31.732(2) Å, b = 10.7145(7) Å, c = 17.6302(14) Å, β = 116.019(3), V = 5386.5(7) Å3, Z = 8, ρcalcd = 1.689 g cm−3, μ = 1.520 mm−1, F(000) = 2768, R1 = 0.0409, wR2 = 0.0772, 3120 independent reflections [θ ≤ 24.999°] and 283 parameters.
Crystal data for compound (3): C24H37BRu2S6, Mr = 730.84, orthorhombic, space group Pbcn, a = 34.1570(11) Å, b = 8.5558(3) Å, c = 19.8431(8) Å, V = 5799.0(4) Å3, Z = 8, ρcalcd = 1.674 g cm−3, μ = 1.487 mm−1, F(000) = 2960, R1 = 0.0457, wR2 = 0.0884, 2850 independent reflections [θ ≤ 24.93°] and 309 parameters.
Crystal data for compound (5): C21H37B3Ru2S2, Mr = 588.19, monoclinic, space group P21/n, a = 8.5681(2) Å, b = 39.1432(9) Å, c = 15.1808(3) Å, β = 95.9220(10), V = 5064.21(19) Å3, Z = 8, ρcalcd = 1.543 g cm−3, μ = 1.363 mm−1, F(000) = 2384, R1 = 0.0420, wR2 = 0.1018, 6613 independent reflections [θ ≤ 23.02°] and 580 parameters.

4. Conclusions

The present work describes the synthesis of various borinane complexes of a group-8 heavier transition metal (i.e., ruthenium) from a dithioformato stabilized arachno-diruthenium pentaborane cluster. The new molecules have similar structures, but they differ in terms of the boron atom’s position in the central six-membered ring {C2S3B}. With a single boron atom in the six-membered ring {C2S3B}, these mono-borinanes can be called 1,3,5-trithia-4-borinane and 1,3,5-trithia-2-borinane complexes of ruthenium. In all the mono-borinane complexes, the six-membered ring {C2BS3} adopt a boat confirmation, which is in contrast to our previously reported trithia-diborinane complexes of ruthenium, [(Cp*Ru){(η2-SCHS)CH2S2(BH2)2}], which adopt a chair conformation. The method reported in this article describing the synthesis of trithia-borinane complexes is unique and may be further utilized to introduce one or more boron atoms to the six-membered ring {C2BS3}. The isolation of these complexes opens up a gateway for the synthesis of early and late transition metal trithia-borinane complexes. Furthermore, in an attempt to convert arachno-[(Cp*Ru)2(B3H8)(CS2H)], 1, to a closo or nido geometry, we performed the pyrolysis of 1 that led to the formation of a capped butterfly cluster. With seven-skeletal-electron-pairs (sep), it satisfies the electron count for a BH capped arachno-butterfly structure. These results demonstrate that both the transition metal and the ligands play an important role in the formation of these complexes. It is interesting to see that the properties and reactivity of molecules can be largely controlled by a variation in the metal or ligand.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/7/2/21/s1. 1H, 11B{1H}, 13C{1H} NMR and mass spectra of compounds 25; The CIF and the checkCIF output files of compounds 2, 3 and 5.

Author Contributions

K.S. and U.K. conceived and designed the experiment; K.S. and U.K. performed the synthesis and the spectroscopic analysis; results were discussed with R.B. and S.G.; R.B. prepared the manuscript with feedback from S.G.; S.G. supervision, S.G. project administration.

Funding

This research was funded by Indo-French Centre for the Promotion of Advanced Research (CEFIPRA), India, grant number 5905-1.

Acknowledgments

DST-FIST, India, is gratefully acknowledged for the HRMS facility. K.S. thank CSIR, India for the research fellowship. We thank V. Ramkumar and P.K. Sudhadevi Antharjanam for X-ray data analysis. X-ray support from Department of Chemistry, IIT Madras and SAIF, IIT Madras, are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hartwig, J.F.; Muhoro, C.N.; He, X.; Eisenstein, O.; Bosque, R.; Maseras, F. Catecholborane Bound to Titanocene. Unusual Coordination of Ligand σ-Bonds. J. Am. Chem. Soc. 1996, 118, 10936–10937. [Google Scholar] [CrossRef]
  2. Douglas, T.M.; Chaplin, A.B.; Weller, A.S. Amine–Borane σ-Complexes of Rhodium. Relevance to the Catalytic Dehydrogenation of Amine–Boranes. J. Am. Chem. Soc. 2008, 130, 14432–14433. [Google Scholar] [CrossRef] [PubMed]
  3. Forster, D.; Tuononen, H.M.; Parvez, M.; Roesler, R. Characterization of β-B-Agostic Isomers in Zirconocene Amidoborane Complexes. J. Am. Chem. Soc. 2009, 131, 6689–6691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tang, C.Y.; Thompson, A.L.; Aldridge, S. Rhodium and Iridium Aminoborane Complexes: Coordination Chemistry of BN Alkene Analogues. Angew. Chem. 2010, 122, 933–937. [Google Scholar] [CrossRef]
  5. Crossley, I.R.; Foreman, M.R.S.J.; Hill, A.F.; White, A.J.P.; Williams, D.J. The first rhodaboratrane: [RhCl(PPh3){B(mt)3}](Rh→B) (mt = methimazolyl). Chem. Commun. 2005, 221–223. [Google Scholar] [CrossRef] [PubMed]
  6. 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]
  7. Greenwood, N.N.; Ward, I.M. Metalloboranes and metal–boron bonding. Chem. Soc. Rev. 1974, 3, 231–271. [Google Scholar] [CrossRef]
  8. Grimes, R.N. Structure and stereochemistry in metalloboron cage compounds. Acc. Chem. Res. 1978, 11, 420–427. [Google Scholar] [CrossRef]
  9. Fehlner, T.P. A molecular orbital analysis of four chromaboranes: On the curious behavior of (η5-C5R5)Cr fragments in a borane cluster environment. J. Organomet. Chem. 1998, 550, 21. [Google Scholar] [CrossRef]
  10. Ghosh, S.; Beatty, A.M.; Fehlner, T.P. The Reaction of Cp*ReH6, Cp* = C5Me5 with Monoborane to Yield a Novel Rhenaborane. Synthesis and Characterization of arachno-Cp*ReH3B3H8. Collect. Czech. Chem. Commun. 2002, 67, 808–812. [Google Scholar] [CrossRef]
  11. Sahoo, S.; Reddy, K.H.K.; Dhayal, R.S.; Mobin, S.M.; Jemmis, E.D.; Ghosh, S. Chlorinated Hypoelectronic Dimetallaborane Clusters Synthesis, Characterization, Electronic Structures of (η5-Cp*W)2B5HnClm (n = 7, m = 2; n = 8, m = 1). Inorg. Chem. 2009, 48, 6509–6516. [Google Scholar] [CrossRef] [PubMed]
  12. 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] [PubMed]
  13. Ghosh, S.; Noll, B.C.; Fehlner, T.P. Expansion of Iridaborane Clusters by Addition of Monoborane. Novel Metallaboranes and Mechanistic Detail. Dalton Trans. 2008, 371–378. [Google Scholar] [CrossRef]
  14. Geetharani, K.; Krishnamoorthy, B.S.; Kahlal, S.; Mobin, S.M.; Halet, J.-F.; Ghosh, S. Synthesis and Characterization of Tantalaboranes. Comparison of the Geometric and Electronic Structures of [(Cp*TaX)2B5H11] (X = Cl, Br and I). Inorg. Chem. 2012, 51, 10176–10184. [Google Scholar] [CrossRef] [PubMed]
  15. Ghosh, S.; Noll, B.C.; Fehlner, T.P. Borane Mimics of Classic Organometallic Compounds: [(Cp*Ru)(B8H14)(RuCp*)]0,+1 Isoelectronic Analogues of Dinuclear Pentalene Complexes. Angew. Chem. Int. Ed. 2005, 44, 6568–6571. [Google Scholar] [CrossRef] [PubMed]
  16. Housecroft, C.E.; Fehlner, T.P. Triborane. A transition metal ligand or heterocluster fragment? Inorg. Chem. 1982, 21, 1739. [Google Scholar] [CrossRef]
  17. Housecroft, C.E. Boranes and Metallaboranes; Ellis Horwood: Chichester, UK, 1990. [Google Scholar]
  18. Mingos, D.M.P. Inorganometallic Chemistry; Fehlner, T.P., Ed.; Plenum: New York, NY, USA, 1992. [Google Scholar]
  19. Hoffmann, R. Building Bridges Between Inorganic and Organic Chemistry (Nobel Lecture). Angew. Chem. Int. Ed. 1982, 21, 711–724. [Google Scholar] [CrossRef]
  20. Chakrahari, K.K.V.; Dudekula, S.; Barik, S.K.; Mondal, B.; Varghese, B.; Ghosh, S. Hypoelectronic Metallaboranes: Synthesis, Structural Characterization, and Electronic Structures of the Metal-Rich Cobaltaboranes. J. Organomet. Chem. 2014, 749, 188–196. [Google Scholar] [CrossRef]
  21. 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, 1483–1487. [Google Scholar] [CrossRef]
  22. Roy, D.K.; Mondal, B.; Shankhari, P.; Anju, R.S.; Geetharani, K.; Mobin, S.M.; Ghosh, S. Supraicosahedral Polyhedra: Synthesis and Structural Characterization of 12, 15 and 16-vertex Rhoda-boranes. Inorg. Chem. 2013, 52, 6705–6712. [Google Scholar] [CrossRef]
  23. Geetharani, K.; Bose, S.K.; Sahoo, S.; Mobin, S.M.; Ghosh, S. Cluster Expansion Reactions of Group 6 and 8 Metallaboranes Using Transition Metal Carbonyl Compounds of Gr 7-9. Inorg. Chem. 2011, 50, 5824–5832. [Google Scholar] [CrossRef] [PubMed]
  24. 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] [PubMed]
  25. Bose, S.K.; Geetharani, K.; Sahoo, S.; Reddy, K.H.K.; Varghese, B.; Jemmis, E.D.; Ghosh, S. Synthesis, Characterization, and Electronic Structure of New Type of Heterometallic Boride Clusters. Inorg. Chem. 2011, 50, 9414–9422. [Google Scholar] [CrossRef] [PubMed]
  26. Anju, R.S.; Roy, D.K.; Mondal, B.; Yuvaraj, K.; Arivazhagan, C.; Saha, K.; Varghese, B.; Ghosh, S. Reactivity of Diruthenium and Dirhodium Analogues of Pentaborane(9): Agostic versus Boratrane Complexes. Angew. Chem. Int. Ed. 2014, 53, 2873–2877. [Google Scholar] [CrossRef] [PubMed]
  27. Saha, K.; Ramalakshmi, R.; Gomosta, S.; Pathak, K.; Dorcet, V.; Roisnel, T.; Halet, J.-F.; Ghosh, S. Design, Synthesis, and Chemistry of Bis(σ)borate and Agostic Complexes of Group 7 Metals. Chem. Eur. J. 2017, 23, 9812–9820. [Google Scholar] [CrossRef] [PubMed]
  28. Saha, K.; Joseph, B.; Borthakur, R.; Ramalakshmi, R.; Roisnel, T.; Ghosh, S. Chemistry of ruthenium σ-borane complex, [Cp*RuCO(μ-H)BH2L] (Cp* = η5-C5Me5; L = C7H4NS2) with terminal and internal alkynes: Structural characterization of vinyl hydroborate and vinyl complexes of ruthenium. Polyhedron 2017, 125, 246–252. [Google Scholar] [CrossRef]
  29. Roy, D.K.; Borthakur, R.; De, A.; Varghese, B.; Phukan, A.K.; Ghosh, S. Synthesis and Characterization of Bis(sigma)borate and Bis-zwitterionic Complexes of Rhodium and Iridium. ChemistrySelect 2016, 1, 3757–3761. [Google Scholar] [CrossRef]
  30. Anju, R.S.; Mondal, B.; Saha, K.; Panja, S.; Varghese, B.; Ghosh, S. Hydroboration of Alkynes with Zwitterionic Ruthenium–Borate Complexes: Novel Vinylborane Complexes. Chem. Eur. J. 2015, 21, 11393–11400. [Google Scholar] [CrossRef]
  31. Ramalakshmi, R.; Saha, K.; Roy, D.K.; Varghese, B.; Phukan, A.K.; Ghosh, S. New Routes to a Series of σ-Borane/Borate Complexes of Molybdenum and Ruthenium. Chem. Eur. J. 2015, 21, 17191–17195. [Google Scholar] [CrossRef]
  32. Anju, R.S.; Roy, D.K.; Geetharani, K.; Mondal, B.; Varghese, B.; Ghosh, S. A fine tuning of metallaborane to bridged-boryl complex, [(Cp*Ru)2(μ-H)(μ-CO)(μ-Bcat)] (cat = 1,2-O2C6H4; Cp* = η5-C5Me5). Dalton Trans. 2013, 42, 12828–12831. [Google Scholar] [CrossRef]
  33. Saha, K.; Kaur, U.; Kar, S.; Mondal, B.; Joseph, B.; Antharjanam, P.K.S.; Ghosh, S. Trithia-diborinane and Bis(bridging-boryl) Complexes of Ruthenium Derived from a [BH3(SCHS)] Ion. Inorg. Chem. 2019. [Google Scholar] [CrossRef] [PubMed]
  34. Sharmila, D.; Yuvaraj, K.; Barik, S.K.; Roy, D.K.; Chakrahari, K.K.; Ramalakshmi, R.; Mondal, B.; Varghese, B.; Ghosh, S. New Heteronuclear Bridged Borylene Complexes That Were Derived from [{Cp*CoCl}2] and Mono-Metal–Carbonyl Fragments. Chem. Eur. J. 2013, 19, 15219–15225. [Google Scholar] [CrossRef] [PubMed]
  35. Bhattacharyya, M.; Prakash, R.; Jagan, R.; Ghosh, S. Synthesis and ligand substitution of tri-metallic triply bridging borylene complexes. J. Organomet. Chem. 2018, 866, 79–86. [Google Scholar] [CrossRef]
  36. Yuvaraj, K.; Bhattacharyya, M.; Prakash, R.; Ramkumar, V.; Ghosh, S. New Trinuclear Complexes of Group 6, 8, and 9 Metals with a Triply Bridging Borylene Ligand. Chem. Eur. J. 2016, 22, 8889–8896. [Google Scholar] [CrossRef]
  37. Bose, S.K.; Roy, D.K.; Shankhari, P.; Yuvaraj, K.; Mondal, B.; Sikder, A.; Ghosh, S. Syntheses and Characterization of New Vinyl-Borylene Complexes by the Hydroboration of Alkynes with [(μ3-BH)(Cp*RuCO)2(μ-CO)Fe(CO)3]. Chem. Eur. J. 2013, 19, 2337–2343. [Google Scholar] [CrossRef]
  38. Yuvaraj, K.; Roy, D.K.; Geetharani, K.; Mondal, B.; Anju, V.P.; Shankhari, P.; Ramkumar, V.; Ghosh, S. Chemistry of Homo- and Heterometallic Bridged-Borylene Complexes. Organometallics 2013, 32, 2705–2712. [Google Scholar] [CrossRef]
  39. Sharmila, D.; Mondal, B.; Ramalakshmi, R.; Kundu, S.; Varghese, B.; Ghosh, S. First-Row Transition-Metal–Diborane and –Borylene Complexes. Chem. Eur. J. 2015, 21, 5074–5083. [Google Scholar] [CrossRef]
  40. Saha, K.; Joseph, B.; Ramalakshmi, R.; Anju, R.S.; Varghese, B.; Ghosh, S. (η4-HBCC-σ,π-Borataallyl Complexes of Ruthenium Comprising an Agostic Interaction. Chem. Eur. J. 2016, 22, 7871–7878. [Google Scholar] [CrossRef]
  41. Bakthavachalam, K.; Yuvaraj, K.; Zafar, M.; Ghosh, S. Reactivity of [M2(μ-Cl)2(cod)2] (M=Ir, Rh) and [Ru(Cl)2(cod)(CH3CN)2] with Na[H2B(bt)2]: Formation of Agostic versus Borate Complexes. Chem. Eur. J. 2016, 22, 17291–17297. [Google Scholar] [CrossRef]
  42. Roy, D.K.; Mondal, B.; Anju, R.S.; Ghosh, S. Chemistry of Diruthenium and Dirhodium Analogues of Pentaborane(9): Synthesis and Characterization of Metal N,S-Heterocyclic Carbene and B-Agostic Complexes. Chem. Eur. J. 2015, 21, 3640–3648. [Google Scholar] [CrossRef]
  43. Saha, K.; Ramalakshmi, R.; Borthakur, R.; Gomosta, S.; Pathak, K.; Dorcet, V.; Roisnel, T.; Halet, J.-F.; Ghosh, S. An Efficient Method for the Synthesis of Boratrane Complexes of Late Transition Metals. Chem. Eur. J. 2017, 23, 18264–18275. [Google Scholar] [CrossRef] [PubMed]
  44. Roy, D.K.; De, A.; Panda, S.; Varghese, B.; Ghosh, S. Chemistry of N,S-Heterocyclic Carbene and Metallaboratrane Complexes: A New η3-BCC-Borataallyl Complex. Chem. Eur. J. 2015, 21, 13732–13738. [Google Scholar] [CrossRef] [PubMed]
  45. Sahoo, S.; Dhayal, R.S.; Varghese, B.; Ghosh, S. Unusual Open Eight-Vertex Oxamolybdaboranes: Structural Characterizations of (η5-C5Me5Mo)2B53-OEt) H6R (R = H and n-BuO). Organometallics 2009, 28, 1586–1589. [Google Scholar] [CrossRef]
  46. Sahoo, S.; Mobin, S.M.; Ghosh, S. Direct Insertion of Sulphur, Selenium and Tellurium atoms into Metallaborane Cages using Chalcogen Powders. J. Organomet. Chem. 2010, 695, 945–949. [Google Scholar] [CrossRef]
  47. Thakur, A.; Sao, S.; Ramkumar, V.; Ghosh, S. Novel Class of Heterometallic Cubane and Boride Clusters Containing Heavier Group 16 Elements. Inorg. Chem. 2012, 51, 8322–8330. [Google Scholar] [CrossRef] [PubMed]
  48. Pandey, K.K. Reactivities of carbonyl sulfide (COS), carbon disulfide (CS2) and carbon dioxide (CO2) with transition metal complexes. Coord. Chem. Rev. 1995, 140, 37–114. [Google Scholar] [CrossRef]
  49. Busetto, L.; Palazzi, A.; Monari, M. Dithiocarbene complexes derived from CS2-bridged dinuclear complexes. J. Organomet. Chem. 1982, 228, C19–C20. [Google Scholar] [CrossRef]
  50. Ramalakshmi, R.; Roisnel, T.; Dorcet, V.; Halet, J.-F.; Ghosh, S. Synthesis and structural characterization of trithiocarbonate complexes of molybdenum and ruthenium derived from CS2 ligand. J. Organomet. Chem. 2017, 849–850, 256–260. [Google Scholar] [CrossRef]
  51. Mondal, B.; Bag, R.; Bakthavachalam, K.; Varghese, B.; Ghosh, S. Synthesis, Structures, and Characterization of Dimeric Neutral Dithiolato-Bridged Tungsten Complexes. Eur. J. Inorg. Chem. 2017, 5434–5441. [Google Scholar] [CrossRef]
  52. Rao, C.E.; Barik, S.K.; Yuvaraj, K.; Bakthavachalam, K.; Roisnel, T.; Dorcet, V.; Halet, J.-F.; Ghosh, S. Reactivity of CS2–Syntheses and Structures of Transition-Metal Species with Dithioformate and Methanedithiolate Ligands. Eur. J. Inorg. Chem. 2016, 4913–4920. [Google Scholar] [CrossRef]
  53. Anju, R.S.; Saha, K.; Mondal, B.; Roisnel, T.; Halet, J.-F.; Ghosh, S. In search for new bonding modes of the methylenedithiolato ligand: novel tri- and tetra-metallic clusters. Dalton Trans. 2015, 44, 11306–11313. [Google Scholar] [CrossRef]
  54. Dallanegra, R.; Chaplin, A.B.; Weller, A.S. Bis(σ-amine–borane) Complexes: An Unusual Binding Mode at a Transition-Metal Center. Angew. Chem. Int. Ed. 2009, 48, 6875–6878. [Google Scholar] [CrossRef]
  55. Marder, T.B.; Lin, Z. (Eds.) Contemporary Metal Boron Chemistry I: Borylenes, Boryls, Borane σ-Complexes, and Borohydrides; Springer-Verlag: Berlin, Germany, 2008; pp. 1–202. [Google Scholar]
  56. Kawano, Y.; Yamaguchi, K.; Miyake, S.; Kakizawa, T.; Shimoi, M. Investigation of the Stability of the M–H–B Bond in Borane σ Complexes [M(CO)51-BH2R⋅L)] and [CpMn(CO)21-BH2R⋅L)] (M = Cr, W; L = Tertiary Amine or Phosphine): Substituent and Lewis Base Effects. Chem. Eur. J. 2007, 13, 6920–6931. [Google Scholar] [CrossRef] [PubMed]
  57. Coffy, T.J.; Medford, G.; Plotkin, J.; Long, G.J.; Huffman, J.C.; Shore, S.G. Metalladiboranes of the iron subgroup: K[M(CO)42-B2H5)] (μ-iron, ruthenium, osmium) and M′(η5-C5H5) (CO)22-B2H5) (M′ = iron, ruthenium). Analogs of metal-olefin complexes). Organometallics 1989, 8, 2404–2409. [Google Scholar] [CrossRef]
  58. Plotkin, J.S.; Shore, S.G. Preparation of (η5-C5H5)(CO)2Fe(η2-B2H5): A neutral metallo-diborane(6) analogue of a metal–olefin complex. J. Organomet. Chem. 1979, 182, C15–C19. [Google Scholar] [CrossRef]
  59. Gloaguen, Y.; Alcaraz, G.; Pécharman, A.-F.; Clot, E.; Vendier, L.; Etienne, S.S. Phosphinoborane and Sulfidoborohydride as Chelating Ligands in Polyhydride Ruthenium Complexes: Agostic σ-Borane versus Dihydroborate Coordination. Angew. Chem. Int. Ed. 2009, 48, 2964–2968. [Google Scholar] [CrossRef]
  60. Hill, A.F.; Owen, G.R.; White, A.J.P.; Williams, D.J. The Sting of the Scorpion: A Metallaboratrane. Angew. Chem. Int. Ed. 1999, 38, 2759–2761. [Google Scholar] [CrossRef]
  61. Bontemps, S.; Gornitzka, H.; Bouhadir, G.; Miqueu, K.; Bourissou, D. Rhodium(I) Complexes of a PBP Ambiphilic Ligand: Evidence for a Metal→Borane Interaction. Angew. Chem. Int. Ed. 2006, 45, 1611–1614. [Google Scholar] [CrossRef]
  62. Figueroa, J.S.; Melnick, J.G.; Parkin, G. Reactivity of the Metal→BX3 Dative σ-Bond:  1,2-Addition Reactions of the Fe→BX3 Moiety of the Ferraboratrane Complex [κ4-B(mimBut)3]Fe(CO)2. Inorg. Chem. 2006, 45, 7056–7058. [Google Scholar] [CrossRef] [PubMed]
  63. Westcott, S.A.; Marder, T.B.; Baker, R.T.; Harlow, R.L.; Calabrese, J.C.; Lam, K.C.; Lin, Z. Reactions of hydroborating reagents with phosphinorhodium hydride complexes: molecular structures of a Rh2B3 metallaborane cluster, an L2Rh(η2-H2BR2) complex and a mixed valence Rh dimer containing a semi-bridging Bcat (cat = 1,2-O2C6H4) group. Polyhedron 2004, 23, 2665–2677. [Google Scholar] [CrossRef]
  64. Braunschweig, H.; Radacki, K.; Rais, D.; Whittell, G.R. A Boryl Bridged Complex: An Unusual Coordination Mode of the BR2 Ligand. Angew. Chem. Int. Ed. 2005, 44, 1192–1194. [Google Scholar] [CrossRef]
  65. Feilong, J.; Fehlner, T.P.; Rheingold, A.L. Preparation of 2,3,4-Tris(η5-cyclopentadienyl)-1,5-diphenyl-1-phospha-2,3,4-tricobaltapentaborane(5); Phenyl Group Migration from Phosphorus to Boron. Angew. Chem. Int. Ed. Engl. 1988, 27, 424–426. [Google Scholar] [CrossRef]
  66. Ibers, J.A. Centenary Lecture. Reactivities of carbon disulphide, carbon dioxide, and carbonyl sulphide towards some transition-metal systems. Chem. Soc. Rev. 1982, 11, 57–73. [Google Scholar] [CrossRef]
  67. Choy, V.J.; O’Connor, C.J. Chelating dioxygen compounds of the platinum metals. Coord. Chem. Rev. 1972, 9, 145–170. [Google Scholar] [CrossRef]
  68. Walther, D. Homogeneous-catalytic reactions of carbon dioxide with unsatureated substrates, reversible CO2-carriers and transcarboxylation reactions. Coord. Chem. Rev. 1987, 79, 135–174. [Google Scholar] [CrossRef]
  69. Anju, R.S.; Saha, K.; Mondal, B.; Dorcet, V.; Roisnel, T.; Halet, J.-F.; Ghosh, S. Chemistry of Diruthenium Analogue of Pentaborane(9) With Heterocumulenes: Toward Novel Trimetallic Cubane-Type Clusters. Inorg. Chem. 2014, 53, 10527–10535. [Google Scholar] [CrossRef] [PubMed]
  70. Coldicott, R.S.; Kennedy, J.D.; Pett, M.T.J. Reactions of carbon disulfide with open nido-6-iridadecaboranes. The formation of closed ten-vertex cluster compounds with boron-to-metal dithioformate bridges and a novel isoclosocloso cluster conversion. J. Chem. Soc. Dalton Trans. 1996, 3819–3824. [Google Scholar] [CrossRef]
  71. Hashimoto, H.; Shang, M.; Fehlner, T.P. Reactions of an Electronically Unsaturated Chromaborane. Coordination of CS2 to (η5-C5Me5)2Cr2B4H8 and Its Hydroboration to a Methanedithiolato Ligand. Organometallics 1996, 15, 1963–1965. [Google Scholar] [CrossRef]
  72. Hartwig, J.F.; Huber, S. Transition metal boryl complexes: structure and reactivity of CpFe(CO)2Bcat and CpFe(CO)2BPh2. J. Am. Chem. Soc. 1993, 115, 4908–4909. [Google Scholar] [CrossRef]
  73. Westcott, A.S.; Marder, T.B.; Baker, R.T. Transition metal-catalyzed addition of catecholborane to α-substituted vinylarenes: hydroboration vs. dehydrogenative borylation. Organometallics 1993, 12, 975–979. [Google Scholar] [CrossRef]
  74. Evans, D.A.; Fu, G.C.; Hoveyda, A.H. Rhodium(I)- and iridium(I)-catalyzed hydroboration reactions: scope and synthetic applications. J. Am. Chem. Soc. 1992, 114, 6671–6679. [Google Scholar] [CrossRef]
  75. Auerhammer, D.; Arrowsmith, M.; Dewhurst, R.D.; Kupfer, T.; Böhnke, J.; Braunschweig, H. Closely related yet different: A borylene and its dimer are non-interconvertible but connected through reactivity. Chem. Sci. 2018, 9, 2252–2260. [Google Scholar] [CrossRef] [PubMed]
  76. Habben, C.; Meller, A.; Noltemeyer, M.; Sheldrick, G.M. Synthese, Molekül- und Kristallstruktur von 3,5-Dimethyl-2,6-bistrimethylsilyl-l-thia-2,4,6-triaza-3,5-diborinan-wolframpentacarbonyl. Z. Naturforsch. 1986, 41b, 799–802. [Google Scholar] [CrossRef]
  77. Matsubara, H.; Tanaka, T.; Takai, Y.; Sawada, M.; Seto, K.; Imazaki, H.; Takahashi, S. Structural Studies of a Liquid Crystalline Compound, 2-(4-Cyanophenyl)-5-(4-butylphenyl)-1,3,2-dioxaborinane, by Means of Nuclear Magnetic Resonance and X-Ray Analyses. Bull. Chem. Soc. Jpn. 1991, 64, 2103–2108. [Google Scholar] [CrossRef] [Green Version]
  78. Slabber, C.A.; Grimmer, C.; Akerman, M.P.; Robinson, R.S. 2-Phenylnaphtho[1,8-de][1,3,2]diazaborinane. Acta Cryst. 2011, E67, o1995. [Google Scholar] [CrossRef] [PubMed]
  79. Wade, K. Structural and Bonding Patterns in Cluster Chemistry. Adv. Inorg. Chem. Radiochem. 1976, 18, 1–66. [Google Scholar] [CrossRef]
  80. Mingos, D.M.P. A General Theory for Cluster and Ring Compounds of the Main Group and Transition Elements. Nat. Phys. Sci. 1972, 236, 99–102. [Google Scholar] [CrossRef]
  81. Mingos, D.M.P. Polyhedral skeletal electron pair approach. Acc. Chem. Res. 1984, 17, 311–319. [Google Scholar] [CrossRef]
  82. Jemmis, E.D.; Balakrishnarajan, M.N.; Pancharatna, P.D. Electronic Requirements for Macropolyhedral Boranes. Chem. Rev. 2002, 102, 93–144. [Google Scholar] [CrossRef]
  83. Ryschkewitsch, G.E.; Nainan, K.C. Octahydrotriborate (1-) ([B3H8]) salts. Inorg. Synth. 1974, 15, 113–114. [Google Scholar] [CrossRef]
  84. Sheldrick, G.M. SHELXS-97; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  85. Sheldrick, G.M. SHELXL; University of Göttingen: Göttingen, Germany, 2014. [Google Scholar]
  86. Altornare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A. Completion and refinement of crystal structures with SIR92. J. Appl. Cryst. 1993, 26, 343–350. [Google Scholar] [CrossRef]
Chart 1. Change in the coordination modes of the molecules with a change in metal or ligand. IV: borane, borate, and diborane; VIX: borane, borate, and agostic; XIXV: metallaboratrane; XVIXX: boryl and borylene complexes.
Chart 1. Change in the coordination modes of the molecules with a change in metal or ligand. IV: borane, borate, and diborane; VIX: borane, borate, and agostic; XIXV: metallaboratrane; XVIXX: boryl and borylene complexes.
Inorganics 07 00021 ch001
Scheme 1. Reaction of [(Cp*Ru)2(B3H8)(CS2H)], 1, in the presence of tellurium powder.
Scheme 1. Reaction of [(Cp*Ru)2(B3H8)(CS2H)], 1, in the presence of tellurium powder.
Inorganics 07 00021 sch001
Figure 1. Molecular structures and labelling diagrams of 2 (a) and 3 (b). Selected bond lengths (Å) and angles (°): 2: B1–S3 1.885(8), S1–Ru1 2.3436(14), Ru1–Ru2 2.7590(6), C21–S5 1.612(5), C23–S3 1.789(6), C23–S2 1.826(5); Ru1–C21–Ru2 86.75(18), S3–C23–S2 118.4(3), S3–B1–S1 119.0(4), Ru1–S1–Ru2 72.36(4). 3: B1–S3 1.922(8), B1–S8 1.879(8), B1–S4 1.968(9), Ru2–S3 2.3364(17), Ru1–Ru2 2.7590(7), C21–S6 1.617(7), C22–S8 1.803(7), C22–S5 1.802(7); Ru2–C21–Ru1 87.2(3), S8–B1–S3 119.7(4), S8–B1–S4 112.5(5), S3–B1–S4 98.9(4), Ru1–S3–Ru2 72.44(5).
Figure 1. Molecular structures and labelling diagrams of 2 (a) and 3 (b). Selected bond lengths (Å) and angles (°): 2: B1–S3 1.885(8), S1–Ru1 2.3436(14), Ru1–Ru2 2.7590(6), C21–S5 1.612(5), C23–S3 1.789(6), C23–S2 1.826(5); Ru1–C21–Ru2 86.75(18), S3–C23–S2 118.4(3), S3–B1–S1 119.0(4), Ru1–S1–Ru2 72.36(4). 3: B1–S3 1.922(8), B1–S8 1.879(8), B1–S4 1.968(9), Ru2–S3 2.3364(17), Ru1–Ru2 2.7590(7), C21–S6 1.617(7), C22–S8 1.803(7), C22–S5 1.802(7); Ru2–C21–Ru1 87.2(3), S8–B1–S3 119.7(4), S8–B1–S4 112.5(5), S3–B1–S4 98.9(4), Ru1–S3–Ru2 72.44(5).
Inorganics 07 00021 g001
Scheme 2. Thermolysis of [(Cp*Ru)2(B3H8)(CS2H)], 1.
Scheme 2. Thermolysis of [(Cp*Ru)2(B3H8)(CS2H)], 1.
Inorganics 07 00021 sch002
Figure 2. Molecular structure and labelling diagram of 5: B1–B2 1.679(11), B1–Ru2 2.098(7), B1–Ru1 2.116(7), B2–Ru2 2.174(6), B2–Ru1 2.231(7), S1–Ru2 2.3017(15), S1–Ru1 2.3035(15), Ru1–Ru2 2.7157(6); Ru2–B1–Ru1 80.2(2), Ru2–B2–Ru1 76.1(2), B1–Ru1–S1 103.2(2), B2–Ru1–S1 83.16(19), B1–Ru1–Ru2 49.59(19), B2–Ru1–Ru2 50.98(16).
Figure 2. Molecular structure and labelling diagram of 5: B1–B2 1.679(11), B1–Ru2 2.098(7), B1–Ru1 2.116(7), B2–Ru2 2.174(6), B2–Ru1 2.231(7), S1–Ru2 2.3017(15), S1–Ru1 2.3035(15), Ru1–Ru2 2.7157(6); Ru2–B1–Ru1 80.2(2), Ru2–B2–Ru1 76.1(2), B1–Ru1–S1 103.2(2), B2–Ru1–S1 83.16(19), B1–Ru1–Ru2 49.59(19), B2–Ru1–Ru2 50.98(16).
Inorganics 07 00021 g002
Table 1. Selected structural and spectroscopic data of borinane derivatives and complexes [33,75,76,77,78].
Table 1. Selected structural and spectroscopic data of borinane derivatives and complexes [33,75,76,77,78].
Entry11B NMR (ppm) adav[B–E] b [Å]Conformations c
Inorganics 07 00021 i0018.3 d1.352half chair
Inorganics 07 00021 i002−5.0 and −15.61.915chair
Inorganics 07 00021 i003f1.414planar
Inorganics 07 00021 i00437.6 1.433 boat
Inorganics 07 00021 i005−11.2 e 1.943 boat
Inorganics 07 00021 i006−4.1 1.919 boat
Inorganics 07 00021 i0077.3 (3)
4.9 (4)
1.923
f
Boat
f
a NMR spectra were recorded in a CDCl3 solvent unless stated. b E = hetero atom in the central ring. c conformation of the central six-membered ring. d In [D6]-acetone. e In CD2Cl2. f Data not available.

Share and Cite

MDPI and ACS Style

Saha, K.; Kaur, U.; Borthakur, R.; Ghosh, S. Synthesis of Trithia-Borinane Complexes Stabilized in Diruthenium Core: [(Cp*Ru)21-S)(η1-CS){(CH2)2S3BR}] (R = H or SMe). Inorganics 2019, 7, 21. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7020021

AMA Style

Saha K, Kaur U, Borthakur R, Ghosh S. Synthesis of Trithia-Borinane Complexes Stabilized in Diruthenium Core: [(Cp*Ru)21-S)(η1-CS){(CH2)2S3BR}] (R = H or SMe). Inorganics. 2019; 7(2):21. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7020021

Chicago/Turabian Style

Saha, Koushik, Urminder Kaur, Rosmita Borthakur, and Sundargopal Ghosh. 2019. "Synthesis of Trithia-Borinane Complexes Stabilized in Diruthenium Core: [(Cp*Ru)21-S)(η1-CS){(CH2)2S3BR}] (R = H or SMe)" Inorganics 7, no. 2: 21. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7020021

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