Next Article in Journal
Utilization of Fe-Ethylenediamine-N,N′-Disuccinic Acid Complex for Electrochemical Co-Catalytic Activation of Peroxymonosulfate under Neutral Initial pH Conditions
Next Article in Special Issue
The Synthesis, Characterization, and Fluxional Behavior of a Hydridorhodatetraborane
Previous Article in Journal
Enrichment Extraction and Activity Study of the Different Varieties of Hericium erinaceus against HCT-8 Colon Cancer Cells
Previous Article in Special Issue
Synthesis, Reactivity and Coordination Chemistry of Group 9 PBP Boryl Pincer Complexes: [(PBP)M(PMe3)n] (M = Co, Rh, Ir; n = 1, 2)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Decaborane: From Alfred Stock and Rocket Fuel Projects to Nowadays †

A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilov Str., 119334 Moscow, Russia
Dedicated to Professor John D. Kennedy on his 80th birthday and in recognition of his outstanding contributions to the chemistry of boranes and metallaboranes.
Molecules 2023, 28(17), 6287; https://doi.org/10.3390/molecules28176287
Submission received: 2 August 2023 / Revised: 21 August 2023 / Accepted: 24 August 2023 / Published: 28 August 2023

Abstract

:
The review covers more than a century of decaborane chemistry from the first synthesis by Alfred Stock to the present day. The main attention is paid to the reactions of the substitution of hydrogen atoms by various atoms and groups with the formation of exo-polyhedral boron–halogen, boron–oxygen, boron–sulfur, boron–nitrogen, boron–phosphorus, and boron–carbon bonds. Particular attention is paid to the chemistry of conjucto-borane anti-[B18H22], whose structure is formed by two decaborane moieties with a common edge, the chemistry of which has been intensively developed in the last decade.

Graphical Abstract

1. Introduction

Decaborane [B10H14] plays a central role in the chemistry of polyhedral boron hydrides. Decaborane is an essential boron reagent for the preparation of medium and higher carboranes C2BnHn+2 (n = 8–10) [1] and the carba-closo-decaborate anions [CB9H10] [2]. Until recently, the synthesis of the closo-dodecaborate [B12H12]2− [3,4] and the carba-closo-dodecaborate [CB11H12] [5,6] anions was also based on the use of decaborane, and it is still used for the synthesis of the closo-decaborate anion [B10H10]2− [7,8]. In addition, decaborane can be used to prepare boron coatings [9,10,11,12,13], nanoparticles [14], microcrystals [15,16], boron nitride nanosheets [17], and nanotubes [18], as well as various metal boride thin films [19,20,21,22,23,24]. Recently, a decaborane-based fuel cell power source with a high energy density was developed [25]. The intensive development of the chemistry of decaborane is associated with the 1950s to the early 1960s, when the main types of its transformations were discovered and described. These early studies were reviewed in the 1960s by Hawthorne [26] and Zakharkin et al. [27]. This area was also partly elucidated in Boron Hydride Chemistry [28] and Comprehensive Inorganic Chemistry I [29]. Recent studies in the field of decaborane chemistry, deepening and expanding the previously described conclusions using modern instrumental methods, were briefly covered in Comprehensive Inorganic Chemistry III [30]. Therefore, the purpose of this review is to give the most complete picture of the current state of the chemistry of decaborane and its derivatives.

2. Synthesis, Structure, and General Properties

The formation of this ten-vertex cluster during the pyrolysis of diborane B2H6 was first described by Alfred Stock and co-workers more than 100 years ago [31,32]. The best yields of decaborane(14) were obtained by heating diborane to 120 °C for 47 h. The low volatility of decaborane allows it to be easily separated from other volatile boron hydrides while being volatile enough to be easily separated from non-volatile products. Decaborane is a colorless, air-stable, easily subliming, malodorous, crystalline solid that melts at 99.7 °C and boils with decomposition at 213 °C [33]. For a long time, interest in the chemistry of boron hydrides was mainly academic but was supported by the fact that boranes and some related compounds did not comply with the usual rules relating the chemical composition to the classical theory of valence. At the same time, various assumptions were made about the structure of B10H14, including linear [34] or naphthalene-like [35] structures.
Practical interest in boron hydrides, and decaborane in particular, arose shortly after World War II, when the United States government launched programs (Projects Hermes, Zip, and HEF (High-Energy Fuels)) [36,37] whose purpose was to develop borane-based aviation and rocket fuels capable of generating much higher energy than conventional kerosene-based fuels [38,39,40]. As a part of this program, two chemical companies, Callery Chemical Company and Olin-Mathieson Corporation, developed eight pilot and production plants and produced an array of borane-derived energetic products, including methyldecaborane (HEF-4), ethyldecaborane (HEF-3), and ethylacetylenedecaborane (HEF-5), to be tested as additives to propellants and explosives [41]. Amost at the same time, due to the development of various physical research methods, such as single-crystal X-ray diffraction, neutron diffraction, and gas phase electron diffraction, the molecular structure of [B10H14] was determined [42,43,44,45,46,47,48,49,50]. The decaborane molecule was found to be shaped like a boat built from ten BH-units, with four additional BHB bridges decorating its bow and stern (Figure 1).
Shortly thereafter, the future Nobel Winner Lipscomb and his collaborators developed bond counting rules and topological principles that made it possible to describe bonding in boron hydrides. According to the topological formalism, the binding in the decaborane molecule can be described by a combination of four 3c-2e B-H-B bonds, six closed or fractional closed 3c-2e B-B-B bonds, and two 2c-2e B-B bonds [51,52,53]. Some time later, molecular orbital theory in the form of the extended Hückel theory was originated and applied to decaborane to provide an alternative to this topological approach [54,55,56]. Subsequently, both the logical basis and the parameters for these molecular orbitals were greatly improved using the more rigorous molecular self-consistent field (SCF) method [57,58]. More recently, the electronic structure of decaborane has been described in terms of the BadeR′s theory “Atoms in Molecules” (AIM) [59].
A powerful tool for determining the structure of polyhedral boron hydrides is NMR spectroscopy, the practical birth of which coincided with a wave of interest in the chemistry of boron hydrides. Therefore, it is not surprising that decaborane was one of the first molecules to be investigated using NMR spectroscopy [60,61,62]. The subsequent development of the instrumental base and methods of NMR spectroscopy caused repeated studies [63,64,65,66,67,68,69,70]. The decaborane molecule has also been characterized by IR [59,71,72], Raman [59], electron [73,74,75], NQR [76,77], photoelectron [78], and electron energy loss [79] spectroscopy. The ionization potentials of decaborane and 11B-enriched decaborane were determined to be 11.0 eV [80] and 10.26 eV [81], respectively. The dipole moment of decaborane was determined by measuring the dielectric constants of benzene, cyclohexane, and carbon disulfide solutions and varies from 3.17 D in carbon disulfide to 3.62 D in benzene [82]. The magnetic susceptibility of decaborane is −116 ± 1.5 × 10−6 emu mol−1 [82]. The heat of formation of decaborane was determined to be −66.1 kJ/mol [83]. The heat capacity of decaborane has been measured, and the derived thermodynamic functions have been calculated [84,85]. The heats of melting and vaporization [85], as well as the vapor pressure of decaborane [85,86], were also determined. Pressure-induced room temperature transformations of decaborane up to 131 GPa were studied using in situ optical spectroscopy techniques [87].
The industrial production of boron hydrides, which involved more than 2000 people, was accompanied by various accidents, which led to the discovery of the high toxicity of decaborane and its derivatives [88,89,90,91,92]. With decaborane, intoxication, headaches, tremors, impaired coordination, confusion, anxiety, photophobia, and other symptoms are observed. Moreover, intoxication can occur from relatively small amounts of decaborane. Decaborane can be detected by its odor at or near its maximum acceptable concentration, but there is considerable olfactory fatigue. Repeated exposure to decaborane can cause severe damage to the nervous system [93]. The effects of decaborane on various animals have also been studied [93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111]. A number of studies were directed to study the mechanism of decaborane action on living organisms [112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130].
The production of decaborane, established in the 1950s, was based on the pyrolytic conversion of diborane proposed by Alfred Stock [33]. At the same time, attempts were made to find an alternative to this dangerous process, among which the use of the CW CO2 laser is worth mentioning [131]. Almost at the same time, a convenient and effective method was proposed, which is based on the oxidation of sodium tetrahydroborate NaBH4 to the octahydrotriborate anion [B3H8], followed by its pyrolysis in diglyme at 105 °C to the tetradecahydro-nido-undecaborate anion [B11H14] [132]. The subsequent mild oxidation of [B11H14] gives decaborane [B10H14] (Scheme 1) [133,134,135,136]. Decaborane can also be obtained by the cage-opening of the closo-decaborate anion [B10H10]2− on protonation with strong acids such as sulfuric acid [137].
Decaborane(14) has an acidic character [138] and can be deprotonated with strong bases, such as sodium hydride [139], tetraalkylammonium hydroxides [140], diethylamine [140], triethylamine [140,141], methylenetriphenylphosphorane [141,142], or a Proton Sponge (PS) [143,144] to give the corresponding salts of the tridecahydro-nido-decaborate [B10H13] anion (Scheme 2). The pKa value of decaborane(14) in aqueous ethanol was found to vary from 2.41 to 3.21 depending on the water content [145].
The (Et3NH)[B10H13] and (Et4N)[B10H13] salts obtained by the deprotonation of decaborane with Et3N and (Et4N)OH, respectively, have been found to trigger the hypergolic reactivity of some polar aprotic organic solvents, such as tetrahydrofuran and ethyl acetate [146].
The solid state structures of (Et3NH)[B10H13] [147], (BnNMe3)[B10H13] [148], and (HPS)[B10H13] [144] were determined by single-crystal X-ray diffraction. The solid state structure of the [B10H13] anion (Figure 2) can be derived from the structure of B10H14 by μ-H(9,10) deprotonation.
In the solution, the [B10H13] anion exists as a mixture of symmetrical and unsymmetrical H-tautomers with different arrangements of bridging hydrogens (Figure 2) [143,149], with an interconversion ΔG value of less than 7 kcal/mol [144].
Strong bases such as sodium hydride in ether solvents are able to remove two protons from decaborane(14) to form the [B10H12]2− dianion [140,150,151]. The latter is unstable in solution and transforms into other decaborates and their derivatives [151]. According to quantum chemical calculations, the [B10H12]2− anion has the C2-symmetric structure with μ-B(5)HB(6) and μ-B(8)HB(9) bridging hydrogens [152].
The reduction of decaborane(14) with KBH4 in water results in the formation of the [arachno-B10H14]2− anion with a boron cage geometry near the same as that of the starting [nido-B10H14] (Scheme 3), which was isolated by precipitation from an aqueous solution in the form of rubidium, cesium, or tetramethylammonium salts [153,154]. The structure of the [B10H14]2− anion was proposed using 11B NMR spectroscopy [155]. The solid state structure of (Me4N)2[B10H14] was determined by single-crystal X-ray diffraction [154]. It was supposed that the reaction proceeds by hydride transfer with the formation of the [B10H15] anion. The latter is unstable in the solution and loses hydrogen to form [nido-B10H13] [156,157].
It should be borne in mind that decaborane itself has pronounced reducing properties. Due to this, the possibility of its use as a reducing agent in organic synthesis was studied. In particular, decaborane can be used for the reduction of acetals to ethers [158], the reductive esterification of aromatic aldehydes [159,160], and the reductive amination of acetals with aromatic amines [161]. Decaborane can also be used for chemoselective reduction aldehydes and ketones [162,163,164], the dehalogenation of α-halocarbonyl compounds [165], and the hydrogenation of alkenes or alkynes [166].
The bridging hydrogens in decaborane(14) were found to exchange rapidly for deuterium atoms with D2O in 1,4-dioxane or acetonitrile to give [μ4-B10H10D4] [72,167,168]. The use of DCl in 1,4-dioxane makes it possible to obtain the decadeuterated decaborane [μ4-5,6,7,8,9,10-B10H4D10] [169]. The treatment of decaborane(14) with DCl in a carbon disulfide solution and in the presence of AlCl3 results in the tetradeuterated decaborane [1,2,3,4-B10H10D4] [169,170]. If the reaction is carried out under heating in sealed ampoule, the product is the octadeuterated decaborane [1,2,3,4,5,7,8,10-B10H6D8] [72]. The tetradeuterated decaborane [5,7,8,10-B10H10D4] was obtained by the reaction of [1,2,3,4,5,7,8,10-B10H6D8] with HCl in a carbon disulfide solution and in the presence of AlCl3 [72]. The octadeuterated decaborane [μ4-1,2,3,4-B10H6D8] was prepared by the reaction of [1,2,3,4-B10H10D4] with D2O in acetonitrile [72], whereas the dodecadeuterated decaborane [μ4-1,2,3,4,5,7,8,10-B10H2D12] was obtained by heating [μ4-B10H10D4] with DCl in carbon disulfide in sealed ampoule in the presence of AlCl3 [72]. Deuterated aromatic solvents can also act as a source of deuterium. For example, heating decaborane(14) in benzene-d6 in the presence of AlCl3 under reflux leads to the the tetradeuterated decaborane [1,2,3,4-B10H10D4], while the reaction of decaborane(14) with AlCl3 in toluene-d8 at 5 °C results in the dideuterated decaborane [2,4-B10H12D2] [171]. The reaction of [1,2,3,4-B10H10D4] with AlCl3 in benzene leads to [1,3-B10H12D2] [172].

3. Halogen Derivatives

Stock first reported the preparation of halogen derivatives of decaborane(14) by the direct reaction of decaborane with halogens in a sealed tube [33]. These reactions were re-investigated in the 1960s. It was found that the reaction of decaborane(14) with 1 equiv. iodine at 110–120 °C leads to the formation of a mixture of 1- and 2-iodo derivatives of decaborane in a ratio of ~1:2 [173,174]. The resulting mixture of isomers can be separated by fractional crystallization from low-boiling alkanes (pentane, hexane, heptane) [173] or chromatographically [175]. The assignment of the substitution position was made based on the 11B NMR spectra [174,176,177]; however, the 1-isomer was initially erroneously assigned as the 5-isomer [173,178]. The subsequent reaction of 2-iododecaborane with iodine at 110 °C results in a mixture of the 1,2- and 2,4-diiodo derivatives [1,2-I2-B10H12] and [2,4-I2-B10H12] in a nearly equal ratio, while the similar reaction of 1-iododecaborane produces mainly [1,2-I2-B10H12]. The reaction of decaborane(14) with an excess of iodine was found to give a mixture of [1,2-I2-B10H12] and [2,4-I2-B10H12] in a ratio of ~1:2 [173]. The reaction of decaborane(14) with iodine or iodine chloride in carbon disulfide in the presence of AlCl3 was found to give a mixture of the 1- and 2-iodo derivatives of decaborane in the same ratio of ~1:2 [179]. Later, the 1- and 2-iodo derivatives of decaborane were synthesized by the reaction of decaborane(14) with iodine chloride in refluxing dichloromethane in the presence of AlCl3 and reliably characterized by NMR spectroscopy [180]. The solid state structures of [1-I-B10H13] [181], [2-I-B10H13] (Figure 3) [180], and [2,4-I2-B10H12] [182] were determined by single-crystal X-ray diffraction.
The 5- and 6-iodo derivatives of decaborane were prepared in an indirect way. The treatment of [arachno-6,9-(Me2S)2-B10H12] with anhydrous HI in benzene under reflux conditions results in a mixture of the 5- and 6-iodo derivatives of decaborane [5-I-B10H13] and [6-I-B10H13] [174], which was separated by fraction crystallization from hexane [174] or chromatographically [175]. The reaction of [arachno-6,9-(Et2S)2-B10H12] with anhydrous HI in benzene at room temperature was found to give the 5-iodo derivative [5-I-B10H13] [183,184,185]. The reaction of (NH4)2[closo-B10H10] with anhydrous HCl in a mixture of AlI3 and 1-butyl-3-methylimidazolium iodide (bmimI) at 70 °C proceeds with the boron cage opening and results in the 6-iodo derivative of decaborane [6-I-B10H13] [186]. The 6-iodo derivative isomerizes to the 5-iodo derivative [5-I-B10H13] in the presence of a catalytic amount of triethylamine in toluene at 60 °C. It is assumed that the isomerization occurs through the transformation of [6-I-B10H13] into the [6-I-B10H13] anion, followed by its isomerization [187]. The 6-iodo derivative [6-I-B10H13] was also found to undergo photochemical isomerization to [5-I-B10H13] under UV-irradiation in a solution [187]. The solid state structures of [5-I-B10H13] [187] and [6-I-B10H13] [186] were determined by single-crystal X-ray diffraction (Figure 4). The 6-iodo derivative [6-I-B10H13] was also obtained by the reaction of (NH4)2[closo-B10H10] with AlI3, followed by the hydrolysis of the resulting intermediate [188].
The reactions of decaborane(14) with bromine in dichloromethane in the presence of AlCl3 [189] or AlBr3 [190], or in carbon disulfide in the presence of AlCl3 [174], lead to the formation of a mixture of 1- and 2-bromo derivatives of decaborane [1-Br-B10H13] and [2-Br-B10H13], which can be separated by fractional crystallization from hexane [173,189,190] or chromatographically [175].
The 5- and 6-bromo derivatives of decaborane were also prepared in an indirect way. The treatment of [arachno-6,9-(Me2S)2-B10H12] with anhydrous HBr in benzene under reflux conditions results in a mixture of the 5- and 6-bromo derivatives of decaborane [5-Br-B10H13] and [6-Br-B10H13] in a ratio of ~1:4, which was separated by fraction crystallization from hexane [174] or chromatographically [175,190]. The reactions of [arachno-6,9-(R2S)2-B10H12] (R = Me, Et) with anhydrous HBr in benzene at room temperature were found to give the 5-bromo derivative [5-Br-B10H13] [183,184,185]. The reaction of (NH4)2[closo-B10H10] with anhydrous HBr in a mixture of AlBr3 and 1-butyl-3-methylimidazolium bromide (bmimBr) at 70 °C proceeds with the boron cage opening and results in the 6-bromo derivative of decaborane [6-Br-B10H13] [186]. The 6-bromo derivative isomerizes to the 5-bromo derivative [5-Br-B10H13] in the presence of a catalytic amount of triethylamine in toluene at 60 °C [187]. The solid state structures of [5-Br-B10H13] [187] and [6-Br-B10H13] [186] were determined by single-crystal X-ray diffraction (Figure 5). The 6-bromo derivative [6-Br-B10H13] was also obtained by the reaction of (NH4)2[closo-B10H10] with AlBr3, followed by the hydrolysis of the resulting intermediate [188].
The reaction of dimethylstannaundecaborane [nido-Me2SnB10H12] with bromine in carbon disulfide leads to the oxidative removal of tin with the formation of the 5,10-dibromo derivative of decaborane [5,10-Br2-B10H12] (Figure 6) [191].
The reaction of decaborane(14) with chlorine in dichloromethane in the presence of AlCl3 results in a mixture of the 1- and 2-chloro derivatives of decaborane [1-Cl-B10H13] and [2-Cl-B10H13] [174,189]. Unexpectedly, a mixture of the 1- and 2-chloro derivatives of decaborane was obtained in the reaction of decaborane(14) with 1,1-difluoroethane in carbon disulfide in the presence of AlCl3 [192]. The isomers were separated by fractional crystallization from pentane or hexane [174,189,192] or chromatographically [175], and the substitution position was assigned using 11B NMR spectroscopy [174,193].
Similar to the corresponding iodo and bromo derivatives, the 5- and 6-chloro derivatives of decaborane were prepared in an indirect way. The treatment of [arachno-6,9-(Me2S)2-B10H12] with anhydrous HCl in benzene under reflux conditions results mainly in the 6-chloro derivative of decaborane [6-Cl-B10H13] with some amount of the 5-chloro isomer [174]. The reactions of [arachno-6,9-(Et2S)2-B10H12] with anhydrous HCl or HgCl2 in benzene at room temperature were also found to give the 6-chloro derivative [6-Cl-B10H13] [183,184,185]. The reaction of (NH4)2[closo-B10H10] with anhydrous HCl in a mixture of AlCl3 and 1-butyl-3-methylimidazolium chloride (bmimCl) at 70 °C proceeds with the boron cage opening and results in the 6-chloro derivative of decaborane [6-Cl-B10H13] [186]. The 6-chloro derivative of decaborane can also be prepared by the reactions of (NH4)2[closo-B10H10] with triflic acid in dichloromethane, respectively [186]. The 6-chloro derivative isomerizes to the 5-chloro derivative [5-Cl-B10H13] in the presence of a catalytic amount of triethylamine in toluene at 60 °C [187]. It was assumed that the isomerization occurs through the transformation of [6-Cl-B10H13] into the [6-Cl-B10H13] anion, followed by its isomerization. The solid state structures of [6-Cl-B10H13], (HPS)[6-Cl-B10H12], and (HPS)[5-Cl-B10H12] were determined by single-crystal X-ray diffraction (Figure 7) [186,187]. The B-H-B bridge deprotonation at the site adjacent to the halogenated boron atoms was revealed [187].
The 6-chloro derivative [6-Cl-B10H13] was also obtained by the reaction of (NH4)2[closo-B10H10] or (Et4N)2[closo-B10H10] with AlCl3, followed by the hydrolysis of the resulting intermediate [188,194,195].
The 6-fluoro derivative of decaborane [6-F-B10H13] was first obtained by the reaction of [arachno-6,9-(Et2S)2-B10H12] with anhydrous HF in benzene at room temperature [183,184,185]. Later, the 6-fluoro derivative was prepared by the reaction of (NH4)2[closo-B10H10] with triflic acid in 1-fluoropentane [186]. The solid state structure of [6-F-B10H13] was determined by single-crystal X-ray diffraction (Figure 8) [186].

4. Derivatives with a B-O Bond

Due to the lack of electrophilic reagents for the introduction of oxygen substituents, the corresponding decaborane derivatives with substituents localized in the “bottom” of the decaborane basket have not yet been obtained. Alkoxy derivatives of decaborane [5-RO-B10H13] (R = Me, Et, Pr, Bu) were first obtained in low (13–20%) yields by trying to iodinate Na[B10H13] in the corresponding esters [196]. The phenoxy derivative [5-PhO-B10H13] was obtained in the same way, using anisole as a solvent [196]. The substitution position was determined using 11B NMR spectroscopy [197]. It was assumed that their formation proceeds through the formation of oxonium derivatives of arachno-decaborane [R2O-B10H13], followed by the elimination of one alkyl group [26]. The reaction of Na[B10H13] with SnCl4 in diethyl ether leads to a mixture of 6- and 5-alkoxy derivatives of decaborane [6-EtO-B10H13] and [5-EtO-B10H13] in a ratio varying from 85:15 to 70:30 depending on the reaction temperature. The isomers were separated using column chromatography on silica [198]. It should be noted that the direct reactions of decaborane(14) with alcohols and phenols ROH leads to its complete degradation to the corresponding trialkyl- or triarylborates (RO)3B [199]. The trimethylsiloxy derivative [6-Me3SiO-B10H13] was obtained in a low yield (5–20%) from the reactions of Na[B10H13] and Na2[B10H12] with Me3SiCl in diethyl ether [198]. The report on the preparation of a trimethylsilyl derivative [Me3Si-B10H13] under similar conditions [200] should apparently be considered erroneous.
The reactions of 5-bromo derivative [5-Br-B10H13] with alcohols ROH in the presence of NaHCO3 in dichloromethane lead to 6-alkoxy derivatives [6-RO-B10H13] (R = Me, Et, t-Bu, c-Hx, CH2CH2SH, CH2CH2I, CH2CH2OCH2CH2Cl, (CH2)3C≡CH, CH(CH2CH=CH2)2), while the reactions of 6-bromo derivative [6-Br-B10H13] with alcohols ROH in the presence of NaHCO3 in dichloromethane lead to 5-alkoxy derivatives [5-RO-B10H13] (R = Me, t-Bu, cHx, CH2CH2SH, CH2CH2I, CH2CH2N(CO)2C2H4, CH2CH2OCH2CH2Cl, (CH2)3C≡CH, CH2C≡CCH3, CH(CH2CH=CH2)2). The reactions of [5-Br-B10H13] and [6-Br-B10H13] with 1,4-cyclohexyldiol lead to the compounds [μ-6,6′-(OC6H10O)-(B10H13)2] and [μ-5,5′-(OC6H10O)-(B10H13)2], respectively. The reactions of alcohols with [6-Br-B10H13] proceed quickly at room temperature, while those with [5-Br-B10H13] require heating (70 °C) to achieve completion. The reaction of [6-Br-B10H13] with ethanol was largely complete after 12 h at room temperature, but the reactions with 2-iodoethanol (~20 h), 2-bromoethanol (~40 h), 2-chloroethanol (~100 h), and 2-fluoroethanol (~125 h) all took increasingly longer times. The reactions with chloro- and iodo-derivatives of decaborane were found to proceed in a similar way; however, the reaction rate decreases in the halogen series I~Br > Cl [198]. The solid state structures of [5-MeO-B10H13] (Figure 9), [6-t-BuO-B10H13] (Figure 9), [5-ClCH2CH2OCH2CH2O-B10H13], [6-ClCH2CH2OCH2CH2O-B10H13], [5-MeC≡CCH2O-B10H13], and [μ-6,6′-(OC6H10O)-(B10H13)2] (Figure 9) were determined by single-crystal X-ray diffraction [201].
The 6-triflate derivative of decaborane [6-TfO-B10H13] was prepared by the reaction of Cs2[closo-B10H10] with neat triflic acid at an ambient temperature [202,203]. In contrast, the reaction of (NH4)2[closo-B10H10] with triflic acid in 1-butyl-3-methylimidazolium triflate at 60 °C results in the 5-triflate derivative of decaborane [5-TfO-B10H13] [204]. It was found that the reaction proceeds through the formation of the 6-triflate derivative, which, upon heating, isomerizes into the 5-triflate derivative. In the presence of a catalytic amount of triethylamine, the isomerization of [6-TfO-B10H13] to [5-TfO-B10H13] proceeds even at room temperature [204]. The reactions of [5-TfO-B10H13] with methanol and 4-methoxyphenol in 1,2-dichloroethane at 70 °C result in the corresponding ethers [6-RO-B10H13] (R = Me, C6H4-4-OMe) [204]. The solid state structures of [6-TfO-B10H13] [202] and [5-TfO-B10H13] [204] were determined by single-crystal X-ray diffraction (Figure 10).
The reactions of Na2[closo-B10H10] with alcohols ROH (R = Me, Et, i-Pr, Bu, Ph) in hexane in the presence of trimethylsilyl triflate lead to the corresponding 6-alkoxy derivatives of decaborane [6-RO-B10H13] [205]. The reaction with water under the same conditions results in the 6-trimethylsiloxy derivative [6-Me3SiO-B10H13] [205].
In a similar way, the reaction of (NH4)2[closo-B10H10] with sulfuric acid produces the 6-hydroxy derivative of decaborane [6-HO-B10H13] [206]. The 6-hydroxy derivative was also obtained as a by-product of the reaction of [arachno-6,9-(Me2S)2-B10H12] with sulfuric acid in benzene [207].
The 6-acetoxy derivative of nido-decaborane [6-AcO-B10H13] was obtained by the reaction of [arachno-6,9-(Me2S)2-B10H12] with mercury acetate [208]. The 6-acetoxy derivative of the arachno-decaborate anion [arachno-6-AcO-B10H13]2− was obtained by the reaction of decaborane with 1-ethyl-3-methylimidazolium acetate (C2mim)(OAc). The solid state structure of (C2mim)2[6-AcO-B10H13] was determined by single-crystal X-ray diffraction (Figure 11) [209].
The bis(decaboranyl) ether [μ-6,6′-O-(B10H13)2] was prepared by the reaction of [arachno-6,9-(R2S)2-B10H12] (R = Me, Et) with sulfuric acid in benzene [210,211]. Its structure was determined by 11B NMR spectroscopy [183,211] and supported by single-crystal X-ray diffraction [212]. The bis(decaboranyl) ether [μ-6,6′-O-(B10H13)2] was also obtained by the dehydration of (H3O)2[closo-B10H10] [213].
The preparation of decaborane derivatives with amides [arachno-6,9-(MeRN(R′)CO)2-B10H12] (R = H, R′ = H, Me; R = Me, R′ = H, Me), triphenylphosphine oxide [arachno-6,9-(Ph3PO)2-B10H12], and dimethylsulfoxide [arachno-6,9-(Me2SO)2-B10H12] has also been reported [214,215,216,217].

5. Derivatives with a B-S Bond

The reaction of decaborane(14) with sulfur in the presence of AlCl3 at 120 °C results in a mixture of the mercapto derivatives [1-HS-B10H13], [2-HS-B10H13], and [1,2-(HS)2-B10H12] [218,219]. The solid state structures of these mercapto derivatives were determined by single-crystal X-ray diffraction (Figure 12) [219].
The reactions of Na2[closo-B10H10] with thiols RSH (R = i-Pr, i-Bu, c-Hx, C6H4-p-Me, C6H4-p-F) in hexane in the presence of trimethylsilyl triflate led to the corresponding 6-alkyl- and 6-arylsulfides [6-RS-B10H13] [205]. It should be noted that the direct reactions of decaborane(14) with alkyl thiols RS lead to its complete degradation to the corresponding trialkylthioborates (RS)3B [220].
Due to its use in the synthesis of carboranes, the 6,9-bis(dimethylsulfonium) derivative of the arachno-decaborate anion [arachno-6,9-(Me2S)2-B10H12], which is formed by refluxing decaborane with dimethyl sulfide in ether or benzene, is the most known decaborane derivative with a B-S bond [214,221,222]. The solid state structure of [arachno-6,9-(Me2S)2-B10H12] was determined by single-crystal X-ray diffraction [223]. The 6,9-bis(dimethylsulfonium) derivative was studied by X-ray and X-ray photoelectron spectroscopy [224,225,226], and its diamagnetic susceptibility was determined [227]. The Me2S substituents in [arachno-6,9-(Me2S)2-B10H12] can be easily replaced by stronger Lewis bases [214,228]. A series of other bis(dialkylsulfonium) derivatives ([arachno-6,9-(RR′S)2-B10H12] (R = R′ = Et, Pr; RR′ = (CH2)4, (CH2CH2)2S, (CH2CH2)2O)) have been prepared in a similar manner [215,220,228,229].
The bis(diethylsulfonium) derivative [arachno-6,9-(Et2S)2-B10H12] can also be obtained by the reaction of (NH4)2[closo-B10H10] with anhydrous hydrogen chloride in diethylsulfide [230]. This approach has been extended to other salts of the closo-decaborate anion and other strong acids, including (H3O)2[closo-B10H10] [231,232].
The reactions of 2-halogen derivatives of decaborane [2-X-B10H13] (X = Cl, Br, I) with dimethylsulfide Me2S give the corresponding 2-halogen-6,9-bis(dimethylsulfonium) derivatives [arachno-2-X-6,9-(Me2S)2-B10H11] [233,234]. In a similar way, the reactions of 5-halogen derivatives of decaborane [5-X-B10H13] (X = F, Br, I) with dialkylsulfides R2S (R = Me, Et) result in the corresponding 5-halogen-6,9-bis(dialkylsulfonium) derivatives [arachno-5-X-6,9-(R2S)2-B10H11], while the reactions of the 6-chloro derivative proceed with the halogen displacement, giving [arachno-6,9-(R2S)2-B10H12] [185]. The solid state structure of [5-Br-6,9-(R2S)2-B10H11] was determined by single-crystal X-ray diffraction [235]. The reactions of 2,4-dichloro- and 1,2,4-trichloro derivatives of decaborane [nido-2,4-Cl2-B10H12] and [nido-1,2,4-Cl3-B10H11] with dimethylsulfide were found to proceed with the boron cage rearrangement, resulting in [arachno-1,7-Cl2-6,9-(Me2S)2-B10H10] and [arachno-1,3,7-Cl3-6,9-(Me2S)2-B10H9], respectively [236]. The solid state structures of [1,7-Cl2-6,9-(Me2S)2-B10H10] and [1,3,7-Cl3-6,9-(Me2S)2-B10H9] were determined by single-crystal X-ray diffraction [236].
The reaction of the 5-triflato derivative of decaborane [5-TfO-B10H13] with dimethylsulfide in toluene results in the 5-triflato-6,9-bis(dialkylsulfonium) derivative [arachno-5-TfO-6,9-B10H11(SMe2)2], while the similar reaction of the 5-triflato derivative [6-TfO-B10H13] proceeds with the substitution of the triflate group, giving [arachno-6,9-B10H12(SMe2)2] [204]. The solid state structure of [5-TfO-6,9-(Me2S)2-B10H11] was determined by single-crystal X-ray diffraction [204].
The 5-dimethylsulfonium derivative of decaborane [nido-5-Me2S-B10H12] was obtained by heating [arachno-6,9-(Me2S)2-B10H12] in toluene or mesitylene, and its solid state structure was determined by single-crystal X-ray diffraction (Figure 13) [215,222,237,238]. It was found that the B-H-B bridge deprotonation occurs at the site adjacent to the substituted boron atom, and thus, the structure of [5-Me2S-B10H12] is similar to the structure of the [6-Cl-B10H12] anion [187].
The preparation of the bis(dimethylthioformamide) [arachno-6,9-(Me2N(H)CS)2-B10H12] [215] and the bis(di(alkyl/aryl)thiourea) [arachno-6,9-((RHN)2CS)2-B10H12] (R = Et, Ph) [217,239] derivatives has also been reported.

6. Derivatives with a B-N Bond

Heating decaborane(14) in acetonitrile under reflux proceeds with hydrogen elimination, resulting in the 6,9-bis(acetonitrilium) derivative of the arachno-decaborate anion [arachno-6,9-(MeC≡N)2-B10H12] [240], which was the first structurally characterized decaborane derivative [217,241,242,243] (Figure 14). The 6,9-bis(acetonitrilium) derivative was studied by X-ray photoelectron spectroscopy [226], and its diamagnetic susceptibility was determined [227]. The 6,9-bis(propionitrilium) and 6,9-bis(benzonitrilium) derivatives [arachno-6,9-(RC≡N)2-B10H12] (R = Et, Ph) were prepared in a similar way from decaborane(14) and propionitrile [244] or benzonitrile [217], respectively. The reaction of the 6,9-bis(acetonitrilium) derivative with diethylcyanamide in diethyl ether results in [arachno-6,9-(Et2NC≡N)2-B10H12] [214,245]. The reaction of the 2-bromo derivative of decaborane [2-Br-B10H13] gives [arachno-6,9-(MeC≡N)2-2-Br-B10H11] [233].
The 6,9-bis(acetonitrilium) derivative [arachno-6,9-(MeC≡N)2-B10H12] reacts with N-nucleophiles (primary and secondary amines and hydrazine), giving the corresponding amidines [arachno-6,9-(RR′N(Me)C=HN)2-B10H12] (R = H, R′ = Et, Pr, Bu, Ph, NH2, NHMe; R = R′ = Et, Pr, Bu] [246,247,248]. The solid state structures of the amidines [6,9-(Bu2N(Me)C=HN)2-B10H12] and [6,9-(PhHN(Me)C=HN)2-B10H12]·Et2O were determined by single-crystal X-ray diffraction (Figure 15) [248]. It should be noted that the reactions with primary amines produce mainly the ZE isomers, whereas the reactions with secondary amines result only in the EE isomers.
The reaction of the 6,9-bis(acetonitrilium) derivative with methanol results in the formation of the corresponding imidate [arachno-6,9-(MeO(Me)C=HN)2-B10H12] [248]. Nowadays, the addition of nucleophiles to the activated triple -C≡N- bond of nitrilium derivatives of various polyhedral boron hydrides has become a widely used method for their modification [249,250,251,252,253,254,255,256,257,258].
The reactions of the 6,9-bis(acetonitrilium) derivative with tertiary amines in refluxing benzene or toluene lead to the corresponding 6,9-bis(trialkylammonium) derivatives [arachno-6,9-(R3N)2-B10H12] (R = Me, Et) [214,245]. The 6,9-bis(ammonium) derivative [arachno-6,9-(H3N)2-B10H12] was prepared by the reaction of decaborane(14) with ammonia in benzene or toluene [259,260]. The solid state structures of [6,9-(H3N)2-B10H12] (Figure 16) [243,261], [6,9-(Me3N)2-B10H12] [262], and [6,9-(Et3N)2-B10H12] [263] were determined by single-crystal X-ray diffraction. [6,9-(H3N)2-B10H12] and [6,9-(Et3N)2-B10H12] were studied by X-ray photoelectron and X-ray fluorescence spectroscopy [225,226,264]. The diamagnetic susceptibilities of [6,9-(Me3N)2-B10H12] and [6,9-(Et3N)2-B10H12] were determined [227]. The thermal decomposition of the 6,9-bis(ammonium) derivative [arachno-6,9-(H3N)2-B10H12] was studied [260,265,266].
The reaction of decaborane(14) with diethylamine in cyclohexane results in the diethylammonium derivative of the arachno-decaborate anion (Et2NH2)[arachno-6-Et2HN-B10H13] [267]. Similar alkylammonium derivatives (Me4N)[arachno-6-RR′R″N-B10H13] (R = Et, R′ = R″ = H; R = R′ = Et, R″ = H; R = R′ = R″ = Et; RR′ = (CH2)5, R″ = H) were prepared by the reactions of Na[B10H13] with the corresponding amines, followed by precipitation with (Me4N)Cl [267]. Heating (Et2NH2)[arachno-6-Et2HN-B10H13] in THF under reflux gives the 6,9-bis(diethylammonium) derivative [arachno-6,9-(Et2HN)2-B10H12], whereas the similar reaction in acetonitrile leads to [arachno-6-Et2HN-9-Et2N(Me)C=HN-B10H12] [267]. The reactions of Na[arachno-6-Et2HN-B10H13] with acetonitrile and dimethylsulfide in the presence of dry HCl result in [arachno-6-Et2NH-9-MeC≡N-B10H12] and [arachno-6-Et2NH-9-Me2S-B10H12], respectively [267]. In a similar way, the reaction of (Me4N)[arachno-6-Et3N-B10H13] with acetonitrile leads to [arachno-6-Et3N-9-MeC≡N-B10H12] [221]. The reactions of Na[arachno-6-Et2HN-B10H13] with amines in THF produce the corresponding 6,9-bis(alkylammonium) derivatives [arachno-6-Et2NH-9-RR′R′’N-B10H12] (R = Et, R′ = R″ = H; R = R′ = Et, R″ = H; R = R′ = R″ = Me) [267].
The reactions of decaborane(14) with pyridines, quinolines, and isoquinoline lead to the corresponding [arachno-6,9-L2-B10H12] (L = pyridine, 2-methylpyridine, 3-methylpyridine, 4-methylpyridine, 2-ethynylpyridine, 2-cyanopyridine, quinoline, 2-methylquinoline, 8-methylquinoline, isoquinoline) derivatives (Scheme 4) [214,236,268,269]. A more convenient way to prepare 6,9-bis(pyridinium) derivatives is the nucleophilic substitution of the dialkylsulfide groups in [arachno-6,9-(R2S)2-B10H12] (R = Me, Et). In this way, the [arachno-6,9-L2-B10H12] (L = pyridine, 2-methylpyridine, 3-methylpyridine, 4-methylpyridine, 2,3-dimethylpyridine, 2,4-dimethylpyridine, 2,5-dimethylpyridine, 2,6-dimethylpyridine, 3,4-dimethylpyridine, 2-methyl-5-ethylpyridine, 2-phenylpyridine, 4-benzylpyridine, 4-styrylpyridine, 2-methoxypyridine, 4-methoxypyridine, 3-chloropyridine, 4-chloropyridine, 2-bromopyridine, 4-bromopyridine, 3,5-dibromopyridine, 3-cyanopyridine, 4-cyanopyridine, 4-acetylpyridine, quinoline) derivatives were synthesized (Scheme 4) [229,270,271,272].
All compounds of this series are brightly colored from yellow to red, which is the reason for the interest in their study by UV and luminescent spectroscopy [229,271,272,273]. The 6,9-bis(pyridinium) derivative [6,9-Py2-B10H12] was studied by X-ray photoelectron and X-ray fluorescence spectroscopy [225,226] and, its diamagnetic susceptibility was determined [227]. The thermal decomposition of the 6,9-bis(pyridinium) and 6,9-bis(quinolinium) derivatives was studied [274]. The solid state structures of [6,9-Py2-B10H12] [275], [6,9-(HC≡C-o-C5H4N)2-B10H12] [269], and [6,9-(N≡C-o-C5H4N)2-B10H12]·CH2Cl2 [269] were determined by single-crystal X-ray diffraction (Figure 17).
It should be noted that the reaction of decaborane(14) with pyridine at low temperatures was found to form the 6,6-bis(pyridinium) derivative [arachno-6,6-Py2-B10H12], which, upon refluxing in dry degassed pyridine, converts into the more stable 6,9-isomer (Scheme 5) [268].
The reaction of [6,9-(Me2S)2-B10H12] with pyrazine in dichloromethane gives the pyrazine-bridged derivative [μ-6,6′-pyrazine-(9-Me2S-B10H12)2], whereas the reaction with 4,4′-bipyridine leads to the product of the substitution of the Me2S groups with azaheterocycle [6,9-(NC5H4C5H4N)2-B10H12] (Figure 18) [276].
The similar reactions of [6,9-(Me2S)2-B10H12] with 1,4-bis[β-(4-pyridyl)vinyl]benzene and 1,4-bis[β-(4-quinolyl)vinyl]benzene were found to produce mixtures of the corresponding bridged and terminal substituted derivatives (Scheme 6) [277].
The reactions of Na[arachno-6-Et2HN-B10H13] and (Me4N)[arachno-6-Et3N-B10H13] with pyridine in THF produce the corresponding 6-alkylammonium-9-pyridinium derivatives [arachno-6-Et2RN-9-Py-B10H12] (R = H, Et) [267,278].
The reactions of decaborane(14) with imidazoles in refluxing benzene result in the corresponding 6,9-bis(imidazolium) derivatives [arachno-6,9-(RIm)2-Me2S-B10H12] (R = H, Me, Et, Bu) (Figure 19). The hypergolic properties of the 6,9-bis(imidazolium) derivatives prepared were studied [279].
The reactions of decaborane(14) with 2-isopropyl- and 2-methyl-5-(2-chloroethyl) tetrazoles in benzene result in the corresponding 6,9-bis(tetrazolium) derivatives [arachno-6,9-L2-Me2S-B10H12] [280].
The 6-isothiocyanato derivative [6-SCN-B10H13] was prepared by the reaction of [6,9-(R2S)2-B10H12] (R = Me, Et) with mercury isothiocyanate [208]. Alternatively, the 6-isothiocyanato derivative can be prepared by the reaction of decaborane(14) with NaSCN in 1,2-dimethoxyethane in the presence of dry HCl [281]. The solid state structure of [6-SCN-B10H13] was determined by single-crystal X-ray diffraction (Figure 20) [281].
The reaction of [6,9-(Me2S)2-B10H12] with excess HN3 in toluene results in the 6-azido-μ-5,6-amino derivative [6-N3-μ-5,6-NH2-B10H11], the structure of which was determined by single-crystal X-ray diffraction (Figure 20) [282].

7. Derivatives with a B-P Bond

The reaction of decaborane(14) with triphenylphosphine in diethyl ether under reflux results in the 6,9-bis(triphenylphosphonium) derivative of the arachno-decaborate anion [arachno-6,9-(Ph3P)2-B10H12]; the same product can be prepared by the reaction of [6,9-(MeC≡N)2-B10H12] with triphenylphosphine in hot acetonitrile [214,245,283]. The solid state structure of [arachno-6,9-(Ph3P)2-B10H12]·2DMF·H2O was determined by single-crystal X-ray diffraction [284]. The 6,9-bis(triphenylphosphonium) derivative was also studied by X-ray emission and X-ray photoelectron spectroscopy [279,280], and its diamagnetic susceptibility was determined [227]. The reaction of [6,9-(Me2S)2-2-Br-B10H11] with triphenylphosphine in benzene results in the corresponding 6,9-bis(triphenylphosphonium) derivative [6,9-(Ph3P)2-2-Br-B10H11] [233].
The reaction of decaborane(14) with PhMe2P at low temperatures (~200 K) gives a mixture of exo,exo- and exo,endo-isomers of [6,9-(PhMe2P)2-arachno-B10H12], which were separated chromatographically [285,286]. The structure of both isomers was confirmed by single-crystal X-ray diffraction (Figure 21) [286].
Under similar conditions, the reaction of the 2,4-dichloro derivatives of decaborane with PhMe2P solely produces the exo,endo-isomer of [6,9-(PhMe2P)2-2,4-Cl2-arachno-B10H10], while the reaction of 2-bromo leads to mixtures of exo,exo- and exo,endo-isomers and [6,9-(PhMe2P)2-2-Br-arachno-B10H11]. It is interesting to note that in the reaction of the 2-bromo derivative, it is precisely the 6,9-exo,endo-isomer that is formed, without any traces of the 9,6-exo,endo-isomer. The solid state structure of the exo,endo-isomer [6,9-(PhMe2P)2-2-Br-arachno-B10H12] was determined by single-crystal X-ray diffraction (Figure 22) [286].
The reaction of decaborane(14) with triethylphosphine in benzene results in the 6,9-bis(triethylphosphonium) derivative [arachno-6,9-(Et3P)2-B10H12] [287], whereas the 6,9-bis(diphenylphosphonium) and 6,9-bis(phenylphosphonium) derivatives [arachno-6,9-(Ph2HP)2-B10H12] and [arachno-6,9-(PhH2P)2-B10H12] were prepared by the reactions of the corresponding phosphines with [arachno-6,9-(Et2S)2-B10H12] [217].
The phosphite, phosphinite, and thiophosphite derivatives of the arachno-decaborate anion [arachno-6,9-(R2R′P)2-B10H12] (R = R′ = OMe, OEt, OPh; R = Ph, R′ = OEt; R = OBu, R′ = Ph; R = R′ = SEt) were prepared by the direct reactions of decaborane(14) with the corresponding phosphorus compounds or via substitution of the Me2S and MeCN groups in [arachno-6,9-(Me2S)2-B10H12] and [arachno-6,9-(MeC≡N)2-B10H12], respectively [217,288,289,290].
The reaction of decaborane(14) with diphenylchlorophosphine in diethyl ether gives the 6,9-bis(chlorodiphenylphosphonium) derivative [arachno-6,9-(ClPh2P)2-B10H12], which, upon the treatment with dimethylamine in alcohols, lead to the corresponding phosphonites [6,9-(ROPh2P)2-B10H12] (R = Me, Et, CH2CH2OH) [290]. The reaction of [6,9-(ClPh2P)2-B10H12] with dimethylamine in aqueous solution water results in the bis(dimethylammonium) salt of the corresponding acid (Me2NH2)2[6,9-(OPh2P)2-B10H12] [290]. The 6,9-bis-(hydroxydiphenylphosphonium) derivative [6,9-(HOPh2P)2-B10H12] was prepared by the reaction of [6,9-(ClPh2P)2-B10H12] with water in acetone [290].
The 6,9-bis(chlorodiphenylphosphonium) derivative reacts with ammonia, hydrazine, primary aliphatic amines, and ethylenimine in alcohols to form the corresponding 6,9-bis(aminodiphenylphosphonium) derivatives [6,9-(R′RNPh2P)2-B10H12] (R = H, R′ =NH2, Me, Bu; R = R′ = CH2CH2) [290]. The reactions of decaborane(14) or [arachno-6,9-(R2S)2-B10H12] (R = Me, Et) with dimethylaminophosphines in refluxing benzene lead to the corresponding 6,9-bis(dimethylaminophosphonium) derivatives [6,9-(RR′(Me2N)P)2-B10H12] (R = R′ = NMe2; R = NMe2, R′ = Ph, Cl; R = R′ = Ph; R = Ph, R′ = Cl; RR′ = OCH2CH2O) [217].
The reaction of [6,9-(ClPh2P)2-B10H12] with NaN3 in ethanol results in the 6,9-bis-(azidodiphenylphosphonium) derivative [6,9-(N3Ph2P)2-B10H12] [290], which, upon the treatment with triphenylphosphine in refluxing benzene, gives the 6,9-bis(triphenylphosphineiminodiphenylphosphonium) derivative [6,9-(Ph3P=NPh2P)2-B10H12] [291].
The bifunctional derivatives [6,9-(XPh2P)2-B10H12] (X = Cl, OH, N3) were used for the synthesis of decaborane-based polymers [291,292]. The chemistry of decaborane-based polymers is considered in detail in the review [293]. The formation of decaborane-based polymers along with a small amount of [6,9-(dppf)2-B10H12] has also been reported in the reaction of decaborane(14) with 1,1′-bis(diphenylphosphino)ferrocene (dppf) [294].
The reaction of Na[B10H13] with tributylphosphine in acetonitrile in the presence of dry HCl leads to [arachno-6-Bu3P-9-MeC≡N-B10H12] [221]. The reaction of Na[B10H13] with diphenylchlorophosphine in diethyl ether leads to the diphenylphosphine derivative [nido-μ-5,6-Ph2P-B10H13] [295], whose structure was determined by single-crystal X-ray diffraction [296]. The same compound was reported to be formed in the reaction of the so-called “Grignard derivative” [B10H13MgI] formed by the treatment of decaborane(14) with MeMgI, with diphenylchlorophosphine in diethyl ether [297]. The diphenylphosphine derivative [nido-μ-5,6-Ph2P-B10H13] is easily deprotonated with triethylamine or sodium hydroxide to form the corresponding salts [295,297]. The solid state structure of the triphenylmethylphosponium salt (Ph3PMe)[arachno-μ-6,9-Ph2P-B10H12] was determined by single-crystal X-ray diffraction (Figure 23) [298].
It should be noted that the reactions of [6,9-(MeC≡N)2-B10H12] with low-coordinated phosphorus compounds, such as phosphaalkynes RC≡P (R = t-Bu, Ad), do not lead to substitution o hydrogens but to the incorporation of phosphorus into the decaborane basket with the formation of 11-vertex phosphoboranes [nido-RC(H)=PB10H13] [299,300].

8. Derivatives with a B-As Bond

The 6,9-bis(trialkyl/arylarsonium) derivatives [6,9-(R3As)2-B10H12] (R = Et, Ph) were prepared by the reactions of decaborane(14) or [6,9-(MeC≡N)2-B10H12] with the corresponding arsines in benzene or toluene [287]. The reaction of decaborane(14) with triethoxyarsine in benzene leads to [6,9-((EtO)3As)2-B10H12] [289].

9. Derivatives with a B-C Bond

Decaborane derivatives with a B-C bond are probably the most studied area of decaborane chemistry. Like the halogenation of decaborane, the direct alkylation reactions result in the substitution of hydrogen atoms at “the bottom” of the decaborane basket. The reaction of decaborane(14) with methyl bromide in carbon disulfide in the presence of AlCl3 at 80 °C gives a mixture of the 2-methyl [2-Me-B10H13], 1,2- and 3,4-dimethyl [1,2-Me2-B10H12] and [2,4-Me2-B10H12], 1,2,3- and 1,2,4-trimethyl [1,2,3-Me3-B10H11] and [1,2,4-Me3-B10H11], 1,2,3,4- and 1,2,3,5(or 8)-tetramethyl [1,2,3,4-Me3-B10H10], and [1,2,3,5(or 8)-Me4-B10H10] derivatives, which were chromatographically separated [301]. The methylation of decaborane(14) was also studied using methyl chloride [302,303].
The reaction of decaborane(14) with neat methyl iodide in the presence of AlCl3 at room temperature gives the 1,2,3,4-tetramethyl derivative [1,2,3,4-Me4-B10H10], whereas the reaction at 120 °C leads to the octasubstituted product [1-I-2,3,4,5,6,7,8-Me7-B10H6]. The similar octasubstituted derivative [1-TfO-2,3,4,5,6,7,8-Me7-B10H6] was obtained by the reaction of decaborane(14) with TfOMe in the presence of a catalytic amount of triflic acid at 120 °C. The solid state structures of [1-I-2,3,4,5,6,7,8-Me7-B10H6] and [1-TfO-2,3,4,5,6,7,8-Me7-B10H6] were determined by single-crystal X-ray diffraction (Figure 24) [304]. The methyl derivatives prepared can be easily deprotonated with a Proton Sponge to give the corresponding salts [304].
The reaction of decaborane(14) with ethyl bromide in carbon disulfide in the presence of AlCl3 under reflux gives a mixture of mono-, di-, and triethyl derivatives [305]. The monoethyl derivative of decaborane was prepared by the reaction of decaborane(14) with neat ethyl bromide in the presence of AlCl3 [306]. The solid state structure of the 1-ethyl derivative of decaborane [1-Et-B10H13] was determined by single-crystal X-ray diffraction [307].
The reaction of decaborane(14) with MeLi in benzene followed by treatment with HCl has been reported to give a mixture of the 6-methyl-, 6,5(or 8)- and 6,9-dimethyl derivatives of decaborane [308]. The reaction with EtLi in benzene gives the 6-ethyl derivative [6-Et-B10H13] [308]. The reaction of decaborane(14) with MeMgI has been shown to proceed by two routes. The major reaction yields the so-called “Grignard derivative” [B10H13MgI] and methane, and the minor reaction produces the 6-methyl derivative of decaborane. The reaction of the “Grignard derivative” with dimethyl sulfate produces a mixture of the 5- and 6-methyl derivatives of decaborane [309]. In a similar way, the reaction of decaborane(14) with EtMgI produces the “Grignard derivative” as the main product and the 6-ethyl derivative of decaborane as a by-product. The reaction of the “Grignard derivative” with [Et3O]BF4 or diethyl sulfate gives the 5-ethyl derivative of decaborane [5-Et-B10H13] [309]. A series of alkyl derivatives [R-B10H13] (R = butyl, amyl, hexyl, cyclohexyl, heptyl, octyl) was prepared by the reactions of the “Grignard derivative” with the corresponding alkyl fluorides [310]. The 6-benzyl derivative of decaborane [6-Bn-B10H13] can be prepared by the reaction of the “Grignard derivative” with benzyl chloride or Na[B10H13] with benzyl bromide [309,311,312].
Later, these reactions were re-examined, and it was shown that the first stage of the reaction of decaborane(14) with the alkyllithium reagents RLi is deprotonation of decaborane with the formation of Li[B10H13]. The reaction with the second equivalent of RLi produces Li2[arachno-6-R-B10H13], which, when treated with HCl, gives Li[arachno-6-R-B10H14] and then [nido-6-R-B10H13] (R = Me, n-Bu, t-Bu). The use of pre-prepared salts of the [B10H13] anion makes it possible to reduce the formation of by-products [237,313].
Another approach to the 6-alkyl derivatives of decaborane includes the reactions of [arachno-6,9-(Me2S)2-B10H12] with alkenes in dichloromethane, resulting in [nido-6-R-8-Me2S-B10H11] (R = cyclohexyl, cyclohexenyl, hexyl, octyl, 2,3-dimethyl-1-butyl, 2,3-dimethyl-2-butyl, 2-methyl-2-butyl, (1R)-(+)-α-pinene, and (1S)-(−)-β-pinene), which can be reduced using Superhydride Li[Et3BH] in THF to [6-R-B10H12] and then protonated with HCl/Et2O to [6-R-B10H13] [222,237,313,314,315]. From the point of view of organic chemistry, these reactions can be considered as the hydroboration reactions. The solid state structures of [6-Chx-8-Me2S-B10H11] [316], [6-Thx-8-Me2S-B10H11] [237] (Figure 25), and [6-Thx-B10H13] [237] (Figure 26) were determined by single-crystal X-ray diffraction.
Another convenient method for the synthesis of 6-alkyl derivatives of decaborane is based on the use of ionic liquids as a solvent. The reactions of decaborane(14) with terminal alkenes in biphasic ionic-liquid/toluene mixtures lead to the corresponding 6-alkyl derivatives [6-R-B10H13] (R = C6H13, C8H17, C16H33, CH(i-Pr)CH2CHMe2, (CH2)2C6H5, (CH2)3C6H5, (CH2)6Br, (CH2)4CH=CH2, (CH2)6CH=CH2, (CH2)3OC3H7, (CH2)3SiMe3, (CH2)4COMe, (CH2)6OAc, (CH2)3OBn, (CH2)3OH, and (CH2)3Bpin, norbornenyl) [317,318,319,320]. The best results were observed for reactions with [bmim]X (1-butyl-3-methylimidazolium, X = Cl or BF4) and bmpyX (1-butyl-4-methylpyridinium, X = Cl or BF4). The reaction mechanism includes the ionic-liquid-promoted formation of the [B10H13] anion, its addition to the alkene to form the [6-R-B10H12] anion, and, finally, the protonation of the last one to form the final product [6-R-B10H13] [314]. The solid state structures of [6-Me3Si(CH2)3-B10H13] and [6-MeC(O)(CH2)4-B10H13] were determined by single-crystal X-ray diffraction (Figure 26) [318].
The 6-cyclohexyl derivative of decaborane [nido-6-C6H11-B10H13] was obtained in a low yield from the reaction of Cs2[closo-B10H10] with triflic acid in cyclohexane (Figure 27) [202,203]. In a similar way, the 6-hexyl derivative [nido-6-C6H13-B10H13] was isolated from the reaction of (NH4)2[closo-B10H10] with concentrated nitric acid in hexane [321].
The Cp2Ti(CO)2-catalyzed reactions of decaborane(14) with terminal alkenes have been found to result in the high-yield formation of 6-alkyl derivatives of decaborane [6-R-B10H13] (R = C6H13, C8H17, (CH2)3SiMe3) [322,323]. The reactions of decaborane(14) with equimolar amounts of bifunctional alkenes such as diallyldimethylsilane, 1,5-hexadiene, 1,4-cyclohexadiene, 1,5-cyclooctadiene, and 2,5-norbornadiene produce the corresponding decaborane derivatives with a double bond in the substituent [6-R-B10H13] (R = (CH2)3SiMe2CH2CH=CH2, (CH2)4CH=CH2, 4-cyclohexenyl, 5-cyclooctenyl, 5-norbornenyl) [322,323,324,325]. The solid state structures of [6-H2C=CHCH2SiMe2(CH2)3-B10H13] (Figure 26) [322], [6-(4′-cyclohexenyl)-B10H13] (Figure 27) [325], and [6-(5′-norbornenyl)-B10H13] (Figure 27) [324] were determined by single-crystal X-ray diffraction.
The reactions of multifunctional alkenes with an excess amount of decaborane(14) produce the saturated linked-cage compounds with two ([μ-6,6′-Me2Si-(6-(CH2)3-B10H13)2], [μ-6,6′-(CH2)6-(B10H13)2], [μ-6,6′-(1″,5″-cyclooctyl)-(B10H13)2], and [μ-6,6′-(2″,5″-norbornyl)-(B10H13)2]) or four ([μ4-6,6′,6″,6‴-Si-(6-(CH2)3-B10H13)4]) decaborane units [322,323,325]. The solid state structures of [μ-6,6′-(CH2)6-(B10H13)2], [μ-6,6′-(2″,5″-norbornyl)-(B10H13)2], and [μ4-6,6′,6″,6‴-Si-(6-(CH2)3-B10H13)4] were determined by single-crystal X-ray diffraction (Figure 28) [322,323,325].
The derivatives with the two decaborane units [μ-6,6′-(1″,5″-cyclooctyl)-(B10H13)2] and [μ-6,6′-(2″,5″-norbornyl)-(B10H13)2] can also be prepared by the titanium-catalyzed reactions of decaborane(14) with [6-(5′-cyclooctenyl)-B10H13] and [6-(5′-norbornenyl)-B10H13], respectively [325].
The 6-alkyl derivatives of decaborane with substituents containing double bonds in the side chain (hexenyl, norbornenyl, etc.) are used for the synthesis of decaborane-based polymers and boron-containing ceramics [319,324,325,326,327,328,329,330,331,332].
The reactions of decaborane(14) with terminal alkenes in the presence of catalytic amounts of PtBr2 or H2PtCl6 lead to the 6,9-dialkyl derivatives nido-[6,9-R2-B10H12] (R = C2H5, C3H7, C4H9, C5H11) [333]. The reactions of [5-TfO-B10H13] and [5-I-B10H13] with 1-pentene in the presence of a catalytic amount of PtBr2 at 55 °C lead to the corresponding 6,9-dialkyl derivatives [6,9-(C5H11)2-5-TfO-B10H11] and [6,9-(C5H11)2-5-I-B10H11] [204]. The solid state structure of [6,9-(C5H11)2-5-I-B10H11] was determined by single-crystal X-ray diffraction (Figure 29) [204].
The reactions of decaborane(14) with terminal alkynes in toluene in the presence of [Cp*IrCl2]2 or [(p-cymene)RuCl2]2 as catalysts lead to the corresponding 6,9-di(β-alkenyl) derivatives of decaborane [6,9-((E)-RCH=CH)2-B10H12] (R = H, C6H13, C6H5, (CH2)2Br, (CH2)3Cl, SiMe3) [334,335]. The solid state structures of [6,9-((E)-Br(CH2)2CH=CH)2-B10H12] and [6,9-((E)-Me3SiCH=CH)2-B10H12] were determined by single-crystal X-ray diffraction (Figure 30) [334,335].
In contrast to [(p-cymene)RuCl2]2, the reactions of decaborane(14) with terminal alkynes in the presence of [(p-cymene)RuI2]2 result in the 6,9-di(α-alkenyl) derivatives [6,9-(R(H2C=)C)2-B10H12] (R = C6H13, CH2-c-C6H11, (CH2)2Br, (CH2)3Cl) [334,335]. The solid state structure of [6,9-(c-C6H11CH2(H2C=)C)2-B10H12] was determined by single-crystal X-ray diffraction (Figure 31) [334,335].
In a similar way, the reactions of 6-alkyldecaboranes [6-R-B10H13] with terminal alkynes in the presence of [Cp*IrCl2]2 give asymmetrically substituted 6-alkyl-9-alkenyl-derivatives [6-R-9-((E)-R′CH=CH)2-B10H12] (R = (CH2)3SiMe3, R′ = H, C6H5, C6H4-m-CH≡CH; CH2CH=CH2; R = C5H11, R′ = H). The solid state structure of [6-Me3Si(CH2)3-9-(E)-m-HC≡CC6H4CH=CH-B10H12] was determined by single-crystal X-ray diffraction (Figure 32) [334,335].
While [Cp*IrCl2]2 proved to be inactive for inducing the hydroboration of simple olefins, such as 1-pentene, by either decaborane or the 6-alkyl-decaboranes, it was found to catalyze the hydroboration of 6-alkyl-9-vinyldecaboranes [6-R-9-CH2=CH-B10H12] (R = C5H11, (CH2)3SiMe3) by 6-alkyl-decaboranes [6-R-B10H13] (R = C5H11, (CH2)3SiMe3) to yield linked-cage products [9,9′-μ-CH2CH2-(6-R-B10H12)2] (R = C5H11, (CH2)3SiMe3) (Figure 33) [335].
The vinyl derivative [6-Me3Si(CH2)3-9-CH2=CH)2-B10H12] was found to readily undergo both homo- and cross-metathesis reactions in the presence of Grubbs’ II catalyst, giving the corresponding products [9,9′-μ-CH=CH-(6-Me3Si(CH2)3-B10H12)2] (Figure 34) and [6-Me3Si(CH2)3-9-RCH=CH-B10H12] (R = C3H7, (CH2)4Br, CH2SiMe3) [335].
Heating [arachno-6,9-(Me2S)2-B10H12] with silylated acetylenes Me3SiC≡CR (R = Me, Bu, SiMe3) leads to the corresponding trimethylsilyl alkenyl derivatives [nido-6-Me3Si(R)C=CH-5-Me2S-B10H11] [336,337]. The solid state structures of [6-Me3Si(Me)C=CH-5-Me2S-B10H11] [336], [6-Me3Si(Bu)C=CH-5-Me2S-B10H11] [337], and [6-(Me3Si)2C=CH-5-Me2S-B10H11] [337] were determined by single-crystal X-ray diffraction (Figure 35).
The reaction of [6,9-(Me2S)2-B10H12] with the phosphaalkyne t-BuC≡P in refluxing benzene leads to [μ-6(C),6′(C),5′(P)-C(t-Bu)PH-(nido-8-Me2S-B10H11)(nido-B10H12)], in which two decaborane units are linked by the C(t-Bu)PH-bridge (Figure 36) [338].
The derivatives [μ-(exo-6(C),endo-6(N)-CH=CH-o-C5H4N)-9(N)-HC≡C-o-C5H4N-arachno-B10H11] (Figure 34) [265] and [μ-(exo-6(C),endo-6(N)-(closo-1′,2′-C2B10H10-2′-)-o-C5H4N)-μ-(exo-8(C),exo-9(N)-CH2CH2-o-C5H4N)-arachno-B10H10] (Figure 37) [339] were isolated in minor amounts as products of the intramolecular hydroboronation of [arachno-6,9-(HC≡C-o-C5H4N)2-B10H12] during its thermolysis in 1,2-dichloroethane. In the latter compound, the formation of the ortho-carborane fragment occurs as a result of the reaction of the acetylene group of the substituent with the second molecule of the decaborane derivative.
The reactions of Cs2[closo-B10H10] with triflic acid or (NH4)2[closo-B10H10] with sulfuric acid in the presence of aromatic hydrocarbons produce the corresponding 6-aryl derivatives of decaborane [nido-6-Ar-B10H13] (Ar = Ph, C6H4-4-Me, C6H3-3,5-Me2, C6H2-2,4,6-Me3, C6H2-2,4,6-iPr3, C6H4Cl, C6H4CF3) [202,203,321,340]. The solid state structures of [6-Ph-B10H13], [6-p-Tol-B10H13], and [6-(2′,4′,6′-iPr3-C6H2-B10H13] were determined by single-crystal X-ray diffraction (Figure 38) [202,203,340].
The 6-phenyl derivative [6-Ph-B10H13] was also obtained by the reaction of decaborane(14) with PhLi, followed by the treatment with HCl in Et2O [237] as well as by the solid state pyrolysis of [nido-Ph2SnB10H12] at 95 °C [198].
The pyrolysis of decaborane(14) in benzene at 200 °C gives the 5-phenyl derivative [5-Ph-B10H13] as the main product together with some amounts of the 6-isomer and 5,8-diphenyl derivative [5,8-Ph2-B10H12] [341]. The solid state structures of [5-Ph-B10H13] and [5,8-Ph2-B10H12] were determined by single-crystal X-ray diffraction (Figure 39) [341].
The pyrolysis of decaborane(14) in toluene at 250 °C affords the novel microporous polymer named “activated borane”, in which the decaborane clusters are interconnected by toluene moieties. Activated borane displays a high surface area of 774 m2 g−1, a thermal stability up to 1000 °C (under Ar), and a sorption capacity to emerging pollutants exceeding the capacity of commercial activated carbon [342].
Heating (HPS)[B10H13] in acetonitrile under reflux results in the formation of the bridged imino derivative (HPS)[arachno-μ-6(C),9(N)-MeC=NH-B10H12] (Figure 40) [343].
The reaction of decaborane(14) with sodium cyanide in water followed by the addition of CsCl gives Cs2[arachno-endo-6-N≡C-B10H13] [215]. The solid state structure of the trimethylphenylammonium salt (Me3NPh)2[endo-6-N≡C-B10H13] (Figure 40) [344] and lead complex {(Bipy)2Pb[endo-6-N≡C-B10H13]} [345] were determined by single-crystal X-ray diffraction.
The reactions of decaborane(14) with sodium cyanide and dimethylsulfide or tetrahydrothiophene lead to the corresponding Na[arachno-6-N≡C-9-RR′S-B10H12] (R = R′ = Me, RR′ = (CH2)4). Na[6-N≡C-9-Me2S-B10H12] can also be prepared by the reaction of [6,9-(Me2S)2-B10H12] with sodium cyanide in dimethylsulfide [215].

10. Derivatives with an exo-Polyhedral B-B Bond

Decaborane derivatives with an exo-polyhedral B-B bond are rare, since the reaction of decaborane(14) with boron hydrides usually leads to the completion of the polyhedral backbone with the formation of tetradecahydro-nido-undecaborate [B11H14] [346] and dodecahydro-closo-dodecaborate [B12H12]2− [3] anions and their derivatives.
The reactions of decaborane(14) with sterically hindered (alkyl/arylimino)(2,2,6,6-tetramethylpiperidino)boranes lead to the corresponding 6-substituted aminoborane derivatives [nido-6-(RNH)(C5H6Me4NH)B-B10H13] (R = t-Bu, C6H3-2,6-iPr2) (Figure 41) [347].
The reaction of Na[B10H13] with 9-bora[3.3.1]bicyclononane (9-Br-BBN) in dichloromethane results in the formation of [μ-5,6-(9-BBN)-B10H13], where the 9-BBN group appears in the role of an asymmetric bridge between the B(5) and B(6) positions of the decaborane basket (Figure 41). It is noteworthy that upon the deprotonation of [μ-5,6-(9-BBN)-B10H13] with a Proton Sponge in dichloromethane, the 9-BBN bridging group migrates from the B(6) atom to the B(10) atom, which leads to the formation of an 11-vertex nido-structure (HPS)[μ-7,7-CH(CH2CH2CH2)2CH-B11H12], being a formal derivative of the [B11H14] anion [348].
Decaborane derivatives with an exo-polyhedral B-B bond also include isomeric conjuncto-decaboranes [B10H13]2, which consist of two nido-B10 units linked by a direct B-B bond. These compounds were first identified as trace impurities in technical decaborane (l4) [349]. In principle, there can be 11 different geometric isomers of conjuncto-decaborane [B10H13]2, 4 of which are in the form of enantiomeric pairs. Therefore, various routes (photolysis [350,351], pyrolysis [350,351], γ-irradiation [352], high-energy electron bombardment [351], silent electrical discharge [353]) for synthesizing these compounds have been developed. All of them, as a rule, lead to the formation of mixtures with different isomeric compositions. The solid state structures of the 1,1′- [353], 1,2′- [354], 1,5′- [352], 1,6′- [238], 2,2′- [355,356], and 2,6′- [356] isomers have been determined by single-crystal X-ray diffraction, whereas the 2,5′- [351], 5,5′- [351], and 6,6′- [357] isomers have been characterized by NMR spectroscopy.
The asymmetric derivatives [arachno-1-(6′-nido-B10H13)-6,9-(Me2S)2-B10H11] and [nido-4-(2′-nido-B10H13)-5-Me2S-B10H11] (Figure 42) were isolated from the reaction of [1,5′-(nido-B10H13)2] with dimethylsulfide under reflux [238]. The first compound was also obtained by the thermolysis of [arachno-6,9-(Me2S)2-B10H12] in refluxing toluene [238].
The isomeric tridecaboranyl species [arachno-1,5-(6′-nido-B10H13)2-6,9-(Me2S)2-B10H10] (Figure 43) and [arachno-1,3-(6′-nido-B10H13)2-6,9-(Me2S)2-B10H10] were isolated from the products of the thermolysis of [arachno-6,9-(Me2S)2-B10H12] in refluxing benzene [238,358].
Here, it is also worth mentioning an unusual structure, in which two hydrogen atoms at positions 5 and 6 of the decaborane basket are replaced by pentaborane moieties—5-(nido-pentaboran-2-yl)-6-(nido-pentaboran-1-yl)-nido-decaborane, formed as a result of the long-term storage (23 years) of a sealed, under-vacuum pentaborane(9) sample under ambient lighting and temperature conditions (Figure 44) [339].

11. Decaborane-Related conjucto-Boranes [B18H22]

Another type of compound worth considering here are the conjuncto-boranes [B18H22], in which two decaborane baskets are connected by a common edge. Octadecaborane(22) [B18H22] in the form of a mixture of syn- and anti-isomers (Figure 45) is formed on the hydrolysis of (H3O)2[trans-B20H18]·nH2O and can be separated by fractional crystallization [359]. More recently, a convenient method has been proposed for the synthesis of anti-[B18H22] by mild oxidation of (Me4N)[nido-B9H12] with iodine in toluene, which gives excellent yields (~80%) and thus provides a large-scale and safe route to this important polyborane cluster [360].
The photophysics of both isomers have been studied by UV–vis spectroscopic techniques and quantum chemical calculations. In an air-saturated hexane solution, anti-[B18H22] shows fluorescence with a high quantum yield, ΦF = 0.97, and singlet oxygen O2(1Δg) production (ΦΔ ~ 0.008). Conversely, the isomer syn-[B18H22] shows no measurable fluorescence, instead displaying a much faster, picosecond nonradiative decay of excited singlet states [361]. Due to this, anti-[B18H22] can be considered as a potential blue laser material. The photophysical properties of anti-[B18H22] can be tuned by the partial substitution of hydrogen atoms with various functional groups. Because of this, combined with its high stability [362,363,364,365,366], anti-[B18H22] is attracting increasing research interest, while the syn-[B18H22] isomer has received much less attention.
The reaction of anti-[B18H22] with chlorine generated in situ from N-chlorosuccinimide (NCS) with HCl/dioxane in dichloromethane leads to the 7-chloro derivative anti-[7-Cl-B18H21] (Figure 46) [367]. The reaction of anti-[B18H22] with AlCl3 in tetrachloromethane results in a mixture of the 3,3′- and 3,4′-dichloro derivatives anti-[3,3′-Cl2-B18H20] (Figure 46) and anti-[3,4′-Cl2-B18H20] (Figure 45) together with minor amounts of the other isomeric dichloro derivatives anti-[4,4′-Cl2-B18H20] (Figure 46), anti-[3,1′-Cl2-B18H20] (Figure 46), and anti-[7,3′-Cl2-B18H20], as well as the 3- and 4-chloro derivatives anti-[3-Cl-B18H21] and anti-[4-Cl-B18H21] and the 3,4,3′- and 3,4,4′-trichloro derivatives anti-[3,4,3′-Cl3-B18H19] and anti-[3,4,4′-Cl3-B18H19], which were separated chromatographically [368].
The bromination of anti-[B18H22] with bromine in dichloromethane in the presence of AlCl3 leads to the 4-bromo or 4,4′-dibromo derivatives anti-[4-Br-B18H21] or anti-[4,4′-Br2-B18H20] depending on the reagent ratio (Figure 47) [369,370].
The reaction of anti-[B18H22] with iodine in ethanol leads to the 4-iodo derivative anti-[7-I-B18H21] (Figure 48) [359,371], while the reaction with I2 or ICl in the presence of AlCl3 in dichloromethane results in the 4-iodo- and 4,4′-diiodo derivatives anti-[4-I2-B18H21] and anti-[4,4′-I2-B18H20] (Figure 48), depending on the reagent ratio [367,371].
The iodine atom in anti-[7-I-B18H21] can be substituted by various nucleophiles: the reaction with trifluoroacetamide in toluene in the presence of K3PO4 gives the corresponding N-boronated amide anti-[7-CF3CONH-B18H21]; the reactions with t-BuOK, 4-FC6H4OK, and 1-AdSK in toluene or tetrahydrofuran lead to the corresponding (thio)ethers anti-[7-RX-B18H21]. The reaction with potassium 2,6-dimethylthiophenolate in toluene results in the corresponding thioether anti-[7-(2′,6′-Me2C6H3S)-B18H21], while the reaction in tetrahydrofuran produces anti-[7-(2′,6′-Me2C6H3S(CH2)4O)-B18H21] [372]. The Pd-catalyzed reactions of anti-[7-I-B18H21] with CF3CONH2, t-BuOK, and 2,6-Me2C6H3OK in the presence of catalytic amounts of RuPhos Pd G4 and RuPhos in 1,4-dioxane lead to the corresponding derivatives with B-N and B-O bonds anti-[7-X-B18H21] [372].
The reaction of anti-[B18H22] with neat methyl iodide in the presence of AlCl3 at room temperature results in the 3,3′,4,4′-tetramethyl derivative anti-[3,3′,4,4′-Me4-B18H18] (Figure 49) as the main product, together with minor amounts of the 4,4′-dimethyl derivative anti-[4,4′-Me2-B18H20] (Figure 49), the 3,4,4′- and 3,3′,4-trimethyl derivatives anti-[3,4,4′-Me3-B18H19] (Figure 47) and anti-[3,3′,4-Me3-B18H19], as well as the 1,3,3′,4,4′- and 3,3′,4,4′,8-pentamethyl derivatives anti-[1,3,3′,4,4′-Me5-B18H17] and anti-[3,3′,4,4′,8-Me5-B18H17] and the 1,1′,3,3′,4,4′-hexamethyl derivative anti-[1,1′,3,3′,4,4′-Me6-B18H16] [373]. The similar reaction with ethyl iodide gives the 3,3′,4,4′-tetraethyl derivative anti-[3,3′,4,4′-Et4-B18H18] (Figure 49) [373].
The dichloroundecamethyl anti-[2,2′-Cl2-1,1′,3,3′,4,4′,7,7′,8,8′,10′-Me11-B18H9], dichlorododecamethyl anti-[2,2′-Cl2-1,1′,3,3′,4,4′,7,7′,8,8′,10,10′-Me12-B18H8] (Figure 50), and dichlorotridecamethyl anti-[2,2′-Cl2-1,1′,3,3′,4,4′,7,7′,8,8′,9,10,10′-Me13-B18H7] derivatives were obtained by the reaction of anti-[B18H22] with methyl iodide in the presence of AlCl3 in dichloromethane at 55 °C [373,374].
The reaction of anti-[B18H22] with elemental sulfur in the presence of AlCl3 at 125 °C leads to the 4,4′-dimercapto derivative anti-[4,4′-(HS)2-B18H20] (Figure 51) [375].
The reaction of anti-[B18H22] with pyridine in refluxing chloroform or benzene unexpectedly results in a two-fold substitution in one of the B10-baskets to form nido-arachno-[6′,9′-Py2-B18H20] (Figure 52) together with some amount of anti-[8′-Py-B18H21] (Figure 52) and [3′,8′-Py2-B16H18] as the main degradation product. In contrast to the thermochromic fluorescence of nido-arachno-[6′,9′-Py2-B18H20] (from 620 nm brick red at room temperature to 585 nm yellow at 8 K), anti-[8′-Py-B18H21] exhibits no luminescence [376,377]. The 6′,9′-disubstituted derivatives with 4-picoline [377], isoquinoline [378], and 5-hydroxyisoquinoline [379] were prepared in a similar way.
The reaction of anti-[B18H22] with methyl isonitrile MeNC in benzene leads to anti-[7-{(MeNH)C3N2HMe2}-B18H20], in which a reductive trimerization of MeNC gives an unusual imidazole-based carbene, {(MeNH)C3N2HMe2}, that is stabilized by coordination to the macropolyhedral boron cluster (Figure 53) [380].
The reaction with tert-butyl isonitrile in 1,2-dichloroethane results in anti-[7-{(t-BuNHCH){t-BuNHC(CN)}CH2}-B18H20], in which a reductive oligomerization of t-BuNC has given the complex polynitrogen base {(t-BuNHCH){t-BuNHC(CN)}CH2:} formally as a zwitterionic carbene attached to the macropolyhedral boron cluster (Figure 54) [381].
The synthesis of few substituted derivatives of syn-[B18H22] was reported. Heating syn-[B18H22] with sulfur in the presence of anhydrous AlCl3 at 125 °C results in a mixture of the isomeric mercapto derivatives syn-[1-HS-B18H21], syn-[3-HS-B18H21], and syn-[4-HS-B18H21], which were all separated by chromatography on silica (Figure 55) [382].
The 3- and 4-mercapto derivatives of syn-[B18H22], syn-[3-HS-B18H21], and syn-[4-HS-B18H21] were obtained as byproducts of thermolysis of nonathiaborane arachno-[SB8H12] in boiling cyclohexane [383]. The mercapto derivatives obtained are brightly luminescent under UV irradiation, making these compounds rare examples of a luminescent derivative of syn-[B18H22] [382,383].
The deprotonation of syn-[B18H22] with NaH in 1,2-dimethoxyethane, followed by the reaction with iodine and dimethylsulfide under reflux, results in the 7-dimethylsulfonium derivative syn-[7-Me2S-B18H20] (Figure 56) [384].

12. Conclusions

The purpose of this review was to give the most complete picture of the current state of the chemistry of decaborane and its derivatives. After its rapid development on the verge of the 1950s and 1960s, associated with the study of the chemistry of decaborane as a potential component of rocket fuels, the chemistry of decaborane was studied by many research groups. However, no comprehensive review elucidating it in full has appeared for more than 50 years, which certainly hindered the development of this important area of boron cluster chemistry. We would like to hope that this review will be useful both for young researchers just starting their way in boron chemistry and for researchers actively working in this field.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation (075-03-2023-642).

Data Availability Statement

No new data were created.

Conflicts of Interest

The author declares no conflict of interests.

Sample Availability

Not applicable.

References

  1. Grimes, R.N. Carboranes, 3rd ed.; Academic Press: London, UK, 2016. [Google Scholar] [CrossRef]
  2. Shmal’ko, A.V.; Sivaev, I.B. Chemistry of carba-closo-decaborate anions [CB9H10]. Russ. J. Inorg. Chem. 2019, 64, 1726–1749. [Google Scholar] [CrossRef]
  3. Sivaev, I.B.; Bregadze, V.I.; Sjöberg, S. Chemistry of closo-dodecaborate anion [B12H12]2−: A review. Collect. Czech. Chem. Commun. 2002, 67, 679–727. [Google Scholar] [CrossRef]
  4. Zhao, X.; Yang, Z.; Chen, H.; Wang, Z.; Zhou, X.; Zhang, H. Progress in three-dimensional aromatic-like closo-dodecaborate. Coord. Chem. Rev. 2021, 444, 214042. [Google Scholar] [CrossRef]
  5. Körbe, S.; Schreiber, P.J.; Michl, J. Chemistry of the carba-closo-dodecaborate(-) anion, CB11H12. Chem. Rev. 2006, 106, 5208–5249. [Google Scholar] [CrossRef]
  6. Douvris, C.; Michl, J. Update 1 of: Chemistry of the carba-closo-dodecaborate(-) anion, CB11H12. Chem. Rev. 2013, 113, R179–R233. [Google Scholar] [CrossRef]
  7. Sivaev, I.B.; Prikaznov, A.V.; Naoufal, D. Fifty years of the closo-decaborate anion chemistry. Collect. Czech. Chem. Commun. 2010, 75, 1149–1199. [Google Scholar] [CrossRef]
  8. Mahfouz, N.; Abi Ghaida, F.; El Hajj, Z.; Diab, M.; Floquet, S.; Mehdi, A.; Naoufal, D. Recent achievements on functionalization within closo-decahydrodecaborate [B10H10]2− clusters. ChemistrySelect 2022, 7, e202200770. [Google Scholar] [CrossRef]
  9. Nakamura, K. Preparation and properties of amorphous boron films deposited by pyrolysis of decaborane in the molecular flow region. J. Electrochem. Soc. 1984, 131, 2691. [Google Scholar] [CrossRef]
  10. Kim, Y.G.; Dowben, P.A.; Spencer, J.T.; Ramseyer, G.O. Chemical vapor deposition of boron and boron nitride from decaborane(14). J. Vac. Sci. Technol. A 1989, 7, 2796–2799. [Google Scholar] [CrossRef]
  11. Perkins, F.K.; Rosenberg, R.A.; Lee, S.; Dowben, P.A. Synchrotron-radiation-induced deposition of boron and boron carbide films from boranes and carboranes: Decaborane. J. Appl. Phys. 1991, 69, 4103–4109. [Google Scholar] [CrossRef]
  12. Saidoh, M.; Ogiwara, N.; Shimada, M.; Arai, T.; Hiratsuka, H.; Koike, T.; Shimizu, M.; Ninomiya, H.; Nakamura, H.; Jimbou, R. Initial boronization of JT-60U tokamak using decaborane. Jpn. J. Appl. Phys. 1993, 32, 3276–3281. [Google Scholar] [CrossRef]
  13. Kodama, H.; Oyaidzu, M.; Sasaki, M.; Kimura, H.; Morimoto, Y.; Oya, Y.; Matsuyama, M.; Sagara, A.; Noda, N.; Okuno, K. Studies on structural and chemical characterization for boron coating films deposited by PCVD. J. Nucl. Mater. 2004, 329–333, 889–893. [Google Scholar] [CrossRef]
  14. Bellott, B.J.; Noh, W.; Nuzzo, R.G.; Girolami, G.S. Nanoenergetic materials: Boron nanoparticles from the pyrolysis of decaborane and their functionalisation. Chem. Commun. 2009, 3214–3215. [Google Scholar] [CrossRef]
  15. Ekimov, E.A.; Zibrov, I.P.; Zoteev, A.V. Preparation of boron microcrystals via high-pressure, high-temperature pyrolysis of decaborane, B10H14. Inorg. Mater. 2011, 47, 1194–1198. [Google Scholar] [CrossRef]
  16. Ekimov, E.A.; Lebed, J.B.; Sidorov, V.A.; Lyapin, S.G. High-pressure synthesis of crystalline boron in B-H system. Sci. Technol. Adv. Mater. 2011, 12, 055009. [Google Scholar] [CrossRef] [PubMed]
  17. Chatterjee, S.; Luo, Z.; Acerce, M.; Yates, D.M.; Johnson, A.T.C.; Sneddon, L.G. Chemical vapor deposition of boron nitride nanosheets on metallic substrates via decaborane/ammonia reactions. Chem. Mater. 2011, 23, 4414–4416. [Google Scholar] [CrossRef]
  18. Chatterjee, S.; Kim, M.J.; Zakharov, D.N.; Kim, S.M.; Stach, E.A.; Maruyama, B.; Sneddon, L.G. Syntheses of boron nitride nanotubes from borazine and decaborane molecular precursors by catalytic chemical vapor deposition with a floating nickel catalyst. Chem. Mater. 2012, 24, 2872–2879. [Google Scholar] [CrossRef]
  19. Kher, S.S.; Spencer, J.T. Chemical vapor deposition precursor chemistry. 3. Formation and characterization of crystalline nickel boride thin films from the cluster-assisted deposition of polyhedral borane compounds. Chem. Mater. 1992, 4, 538–544. [Google Scholar] [CrossRef]
  20. Kher, S.S.; Tan, Y.; Spencer, J.T. Chemical vapor deposition of metal borides. 4. The application of polyhedral boron clusters to the chemical vapor deposition formation of gadolinium boride thin-film materials. Appl. Organomet. Chem. 1996, 10, 297–304. [Google Scholar] [CrossRef]
  21. Kher, S.S.; Romero, J.V.; Caruso, J.D.; Spencer, J.T. Chemical vapor deposition of metal borides. 6. The formation of neodymium boride thin film materials from polyhedral boron clusters and metal halides by chemical vapor deposition. Appl. Organomet. Chem. 2008, 22, 300–307. [Google Scholar] [CrossRef]
  22. Tynell, T.; Aizawa, T.; Ohkubo, I.; Nakamura, K.; Mori, T. Deposition of thermoelectric strontium hexaboride thin films by a low pressure CVD method. J. Cryst. Growth 2016, 449, 10–14. [Google Scholar] [CrossRef]
  23. Guélou, G.; Martirossyan, M.; Ogata, K.; Ohkubo, I.; Kakefuda, Y.; Kawamoto, N.; Kitagawa, Y.; Ueda, J.; Tanabe, S.; Maeda, K.; et al. Rapid deposition and thermoelectric properties of ytterbium boride thin films using hybrid physical chemical vapor deposition. Materialia 2018, 1, 244–248. [Google Scholar] [CrossRef]
  24. Ohkubo, I.; Aizawa, T.; Nakamura, K.; Mori, T. Control of competing thermodynamics and kinetics in vapor phase thin-film growth of nitrides and borides. Front. Chem. 2021, 9, 642388. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, Y.; Peng, Y.; Wan, Q.; Ye, D.; Wang, A.; Zhang, L.; Jiang, W.; Liu, Y.; Li, J.; Zhuang, X.; et al. Fuel cell power source based on decaborane with high energy density and low crossover. Mater. Today Energy 2023, 32, 101244. [Google Scholar] [CrossRef]
  26. Hawthorne, M.F. Decaborane-14 and its derivatives. In Advances in Inorganic Chemistry and Radiochemistry; Academic Press: New York, NY, USA, 1963; Volume 5, pp. 307–347. [Google Scholar] [CrossRef]
  27. Stanko, V.I.; Chapovskii, Y.A.; Brattsev, V.A.; Zakharkin, L.I. The chemistry of decaborane and its derivatives. Russ. Chem. Rev. 1965, 34, 424–439. [Google Scholar] [CrossRef]
  28. Shore, S.G. Nido- and arachno-boron hydrides. In Boron Hydride Chemistry; Muetterties, E.L., Ed.; Academic Press: New York, NY, USA, 1975; pp. 79–174. [Google Scholar] [CrossRef]
  29. Greenwood, N.N. Boron. In Comprehensive Inorganic Chemistry; Pergamon Press: Oxford, UK, 1973; Volume 1, pp. 818–837. [Google Scholar] [CrossRef]
  30. Sivaev, I.B. Molecular boron clusters. In Comprehensive Inorganic Chemistry; Elsevier: Amsterdam, The Netherlands, 2023; Volume 1, pp. 740–777. [Google Scholar] [CrossRef]
  31. Stock, A.; Friederici, K.; Priess, O. Borwasserstoffe. III. Feste Borwasserstoffe: Zur Kenntnis des B2H6. Ber. Dtsch. Chem. Ges. 1913, 46, 3353–3365. [Google Scholar] [CrossRef]
  32. Stock, A.; Pohland, E. Borwasserstoffe. XII. Zur Kenntnis des B10H14. Ber. Dtsch. Chem. Ges. 1929, 62, 90–99. [Google Scholar] [CrossRef]
  33. Stock, A. Hydrides of Boron and Silicon; Cornell University Press: Ithaca, NY, USA, 1933. [Google Scholar]
  34. Stock, A. Borwasserstoffe. XI. Strukturformeln der Borhydride. Ber. Dtsch. Chem. Ges. 1926, 59, 2226–2229. [Google Scholar] [CrossRef]
  35. Huggins, M.L. Boron hydrides. J. Phys. Chem. 1922, 26, 833–835. [Google Scholar] [CrossRef]
  36. Clark, J.D. Ignition! An Informal History of Liquid Rocket Propellants; Rutgers University Press: New Brunswick, NJ, USA, 1972. [Google Scholar] [CrossRef]
  37. Boron High Energy Fuels: Hearings Before the Committee on Science and Astronautics U.S. House of Representatives. Eighty-Six Congress, First Session; United States Government Printing Office: Washington, DC, USA, 1959.
  38. Goodger, E. Unconventional fuels. New Sci. 1957, 27–29. [Google Scholar]
  39. Siegel, B.; Mack, J.L. The boron hydrides. J. Chem. Ed. 1957, 34, 314–317. [Google Scholar] [CrossRef]
  40. Martin, D.R. The development of borane fuels. J. Chem. Ed. 1959, 36, 208–214. [Google Scholar] [CrossRef]
  41. Gold, J.W.; Militscher, C.; Slauson, D.D. Pentaborane Disposal: Taming the Dragon. Available online: https://dokumen.tips/documents/pentaborane-taming-the-dragonpdf.html (accessed on 23 August 2023).
  42. Kasper, J.S.; Lucht, C.M.; Harker, D. The structure of the decaborane molecule. J. Am. Chem. Soc. 1948, 70, 881. [Google Scholar] [CrossRef]
  43. Kasper, J.S.; Lucht, C.M.; Harker, M. The crystal structure of decaborane, B10H14. Acta Cryst. 1950, 3, 436–455. [Google Scholar] [CrossRef]
  44. Moore, E.B.; Dickerson, R.E.; Lipscomb, W.N. Least squares refinements of B10H14, B4H10, and B5H11. J. Chem. Phys. 1957, 27, 209–211. [Google Scholar] [CrossRef]
  45. Brill, R.; Dietrich, H.; Dierks, H. Distribution of the bonding electrons in decaborane-14. Angew. Chem. Int. Ed. 1970, 9, 524–526. [Google Scholar] [CrossRef]
  46. Brill, R.; Dietrich, H.; Dierks, H. Die Verteilung der Bindungselektronen im Dekaboran-molekül (B10H14). Acta Cryst. B 1971, 27, 2003–2018. [Google Scholar] [CrossRef]
  47. Dietrich, H.; Scheringer, C. Refinement of the molecular charge distribution in decaborane(14). Acta Cryst. B 1978, 34, 54–63. [Google Scholar] [CrossRef]
  48. Tippe, A.; Hamilton, W.C. Neutron diffraction study of decaborane. Inorg. Chem. 1969, 8, 464–470. [Google Scholar] [CrossRef]
  49. Vilkov, L.V.; Mastryukov, V.S.; Akishin, P.A. An electron-diffraction study of the structure of the decaborane molecule in the vapor state. J. Struct. Chem. 1964, 4, 301–304. [Google Scholar] [CrossRef]
  50. Mastryukov, V.S.; Dorofeeva, O.V.; Vilkov, L.V. Reexamination of the electron-diffraction data for the decaborane(14) molecule. J. Struct. Chem. 1975, 16, 110–112. [Google Scholar] [CrossRef]
  51. Eberhardt, W.H.; Crawford, B.; Lipscomb, W.N. The valence structure of the boron hydrides. J. Chem. Phys. 1954, 22, 989–1001. [Google Scholar] [CrossRef]
  52. Lipscomb, W.N. Valence in the boron hydrides. J. Phys. Chem. 1957, 61, 23–27. [Google Scholar] [CrossRef]
  53. Lipscomb, W.N. Boron Hydrides; W. A. Benjamin, Inc.: New York, NY, USA, 1963. [Google Scholar]
  54. Moore, E.B.; Lohr, L.L.; Lipscomb, W.N. Molecular orbitals in some boron compounds. J. Chem. Phys. 1961, 35, 1329–1334. [Google Scholar] [CrossRef]
  55. Moore, E.B. Molecular orbitals in B10H14. J. Chem. Phys. 1962, 37, 675–677. [Google Scholar] [CrossRef]
  56. Hoffmann, R.; Lipscomb, W.N. Boron hydrides: LCAO-MO and resonance studies. J. Chem. Phys. 1962, 37, 2872–2883. [Google Scholar] [CrossRef]
  57. Laws, E.A.; Stevens, R.M.; Lipscomb, W.N. Self-consistent field study of decaborane(14). J. Am. Chem. Soc. 1972, 94, 4467–4474. [Google Scholar] [CrossRef]
  58. Lipscomb, W.N. Advances in theoretical studies of boron hydrides and carboranes. In Boron Hydride Chemistry; Muetterties, E.L., Ed.; Academic Press: New York, NY, USA, 1975; pp. 39–78. [Google Scholar] [CrossRef]
  59. Kononova, E.G.; Klemenkova, Z.S. The electronic structure of nido-B10H14 and [6-Ph-nido-6-CB9H11] in terms of Bader’s theory (AIM). J. Mol. Struct. 2013, 1036, 311–317. [Google Scholar] [CrossRef]
  60. Shoolery, J.N. The relation of high resolution nuclear magnetic resonance spectra to molecular structures. Discuss. Faraday Soc. 1955, 19, 215–225. [Google Scholar] [CrossRef]
  61. Williams, R.E.; Shapiro, I. Reinterpretation of nuclear magnetic resonance spectra of decaborane. J. Chem. Phys. 1958, 29, 677–678. [Google Scholar] [CrossRef]
  62. Phillips, W.D.; Miller, H.C.; Muetterties, E.L. B11 Magnetic resonance study of boron compounds. J. Am. Chem. Soc. 1959, 81, 4496–4500. [Google Scholar] [CrossRef]
  63. Pilling, R.L.; Tebbe, F.N.; Hawthorne, M.F.; Pier, E.A. Boron-11 nuclear magnetic resonance spectra of two boron hydride derivatives at 60 Mc./sec. Proc. Chem. Soc. 1964, 402–403. [Google Scholar] [CrossRef]
  64. Keller, P.C.; MacLean, D.; Schaeffer, R. Final assignment of B1 (3) N.M.R. resonance of decaborane. Chem. Commun. 1965, 204. [Google Scholar] [CrossRef]
  65. Williams, R.L.; Greenwood, N.N.; Morris, J.H. The nuclear magnetic resonance spectrum of decaborane-14. Spectrochim. Acta 1965, 21, 1579–1587. [Google Scholar] [CrossRef]
  66. Bodner, G.M.; Sneddon, L.G. An assignment of the hydrogen-1 magnetic resonance spectrum of decaborane at 220 MHz. Inorg. Chem. 1970, 9, 1421–1423. [Google Scholar] [CrossRef]
  67. Onak, T.; Marynick, D. Application of a ring current model to decaborane(14). A correlation of proton and boron-11 nuclear magnetic resonance chemical shifts. Trans. Faraday Soc. 1970, 66, 1843–1847. [Google Scholar] [CrossRef]
  68. Greenwood, N.N.; Kennedy, J.D. N.M.R. evidence for a transition between isotropic and anisotropic thermal motion of nido-decaborane in an aromatic solvent. J. Chem. Soc. Chem. Commun. 1979, 1099–1101. [Google Scholar] [CrossRef]
  69. Colquhoun., I.J.; McFarlane, W. Heteronuclear two-dimensional nuclear magnetic resonance: Decaborane. J. Chem. Soc. Dalton Trans. 1981, 2014–2016. [Google Scholar] [CrossRef]
  70. Venable, T.L.; Hutton, W.C.; Grimes, R.N. Two-dimensional boron-11-boron-11 nuclear magnetic resonance spectroscopy as a probe of polyhedral structure: Application to boron hydrides, carboranes, metallaboranes, and metallacarboranes. J. Am. Chem. Soc. 1984, 106, 29–37. [Google Scholar] [CrossRef]
  71. Keller, W.E.; Johnston, H.L. A note on the vibrational frequencies and the entropy of decaborane. J. Chem. Phys. 1952, 20, 1749–1751. [Google Scholar] [CrossRef]
  72. Hanousek, F.; Štíbr, B.; Heřmánek, S.; Plešek, J. Chemistry of boranes. XXXI. Infrared spectra of decaborane and its deuterio derivatives. Collect. Czech. Chem. Commun. 1973, 38, 1312–1320. [Google Scholar] [CrossRef]
  73. Pimentel, G.C.; Pitzer, K.S. The ultraviolet absorption and luminescence of decaborane. J. Chem. Phys. 1949, 17, 882–884. [Google Scholar] [CrossRef]
  74. Pondy, P.R.; Beachell, H.C. Near infrared spectra of diborane, pentaborane, and decaborane. J. Chem. Phys. 1956, 25, 238–241. [Google Scholar] [CrossRef]
  75. Haaland, A.; Eberhardt, W.H. Electronic spectrum of decaborane. J. Chem. Phys. 1962, 36, 2386–2392. [Google Scholar] [CrossRef]
  76. Lötz, A.; Olliges, J.; Voitländer, J. The nuclear quadrupole double resonance spectrum of decaborane. Chem. Phys. Lett. 1982, 93, 560–563. [Google Scholar] [CrossRef]
  77. Hiyama, Y.; Butler, L.G.; Brown, T.L. Boron-10 and boron-11 nuclear quadrupole resonance spectrum of decaborane[14]. J. Magn. Res. 1985, 65, 472–480. [Google Scholar] [CrossRef]
  78. Lloyd, D.R.; Lynaugh, N.; Roberts, P.J.; Guest, M.F. Photoelectron studies of boron compouds. Part 5. Higher boron hydrides B4H10, B5H9 and B10H14. J. Chem. Soc. Faraday Trans. 2 1975, 71, 1382–1394. [Google Scholar] [CrossRef]
  79. Hitchcock, A.P.; Wen, A.T.; Lee, S.; Glass, J.A.; Spencer, J.T.; Dowben, P.A. Inner-shell excitation of boranes and carboranes. J. Phys. Chem. 1993, 97, 8171–8181. [Google Scholar] [CrossRef]
  80. Margrave, J.L. Ionization potentials of B5H9, B5H8I, B10H14, and B10H13C2H5 from electron impact studies. J. Chem. Phys. 1960, 32, 1889. [Google Scholar] [CrossRef]
  81. Kaufman, J.J.; Koski, W.S.; Kuhns, L.J.; Law, R.W. Appearance and ionization potentials of selected fragments from decaborane, B1110H141. J. Am. Chem. Soc. 1962, 84, 4198–4205. [Google Scholar] [CrossRef]
  82. Bottei, R.S.; Laubengayer, A.W. The dipole moment and magnetic susceptibility of decaborane. J. Phys. Chem. 1962, 66, 1449–1451. [Google Scholar] [CrossRef]
  83. Johnson, W.H.; Kilday, M.V.; Prosen, E.J. Heat of formation of decaborane. J. Res. Natl. Bur. Stand. A Phys. Chem. 1960, 64, 521–525. [Google Scholar] [CrossRef] [PubMed]
  84. Kerr, E.C.; Hallett, N.C.; Johnston, H.L. Low temperature heat capacity of inorganic solids. VI. The heat capacity of decaborane, B10H14, from 14 to 305 K. J. Am. Chem. Soc. 1951, 73, 1117–1119. [Google Scholar] [CrossRef]
  85. Furukawa, G.T.; Park, R.P. Heat capacity, heats of fusion and vaporization, and vapor pressure of decaborane (B10H14). J. Res. Natl. Bur. Stand. 1955, 55, 255–260. [Google Scholar] [CrossRef]
  86. Miller, G.A. The vapor pressure of solid decaborane. J. Phys. Chem. 1963, 67, 1363–1364. [Google Scholar] [CrossRef]
  87. Nakano, S.; Hemley, R.J.; Gregoryanz, E.A.; Goncharov, A.F.; Mao, H.-K. Pressure-induced transformations of molecular boron hydride. J. Phys. Condens. Matter 2002, 14, 10453. [Google Scholar] [CrossRef]
  88. Rozendaal, H.M. Clinical observation on the toxicology of boron hydrides. AMA Arch. Ind. Hyg. Occup. Med. 1951, 4, 257–260. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1951-09_4_3/page/256/mode/2up (accessed on 31 July 2023).
  89. Krackow, E.H. Toxicity and health hazards of boron hydrides. AMA Arch. Ind. Hyg. Occup. Med. 1953, 8, 335–339. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1953-10_8_4/page/334/mode/2up (accessed on 31 July 2023).
  90. Cole, V.V.; Hill, D.L.; Oikemus, A.H. Problems in the study of decaborane and possible therapy of its poisoning. AMA Arch. Ind. Hyg. Occup. Med. 1954, 10, 158–161. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1954-08_10_2/page/158/mode/2up (accessed on 31 July 2023).
  91. Lowe, H.; Freeman, G. Boron hydride (borane) intoxication in man. AMA Arch. Ind. Hyg. Occup. Med. 1957, 12, 523–533. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1957-12_16_6/page/522/mode/2up (accessed on 31 July 2023). [CrossRef]
  92. Cordasco, E.M.; Cooper, R.W.; Murphy, J.V.; Anderson, C. Pulmonary aspects of some toxic experimental space fuels. Dis. Chest 1962, 41, 68–72. [Google Scholar] [CrossRef]
  93. Merritt, J.H. Pharmacology and toxicology of propellant fuels: Boron hydrides. Aeromed. Rev. 1966, 3, 1–11. [Google Scholar]
  94. Svirbely, J.L. Acute toxicity studies of decaborane and pentaborane by inhalation. AMA Arch. Ind. Hyg. Occup. Med. 1954, 10, 298–304. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1954-10_10_4/page/298/mode/2up (accessed on 31 July 2023).
  95. Svirbely, J.L. Subacute toxicity of decaborane and pentaborane vapors. AMA Arch. Ind. Hyg. Occup. Med. 1954, 10, 305–311. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1954-10_10_4/page/304/mode/2up (accessed on 31 July 2023).
  96. Svirbely, J.L. Toxicity tests of decaborane for laboratory animals. I. Acute toxicity studies. AMA Arch. Ind. Hyg. Occup. Med. 1955, 11, 132–137. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1955-02_11_2/page/132/mode/2up (accessed on 31 July 2023).
  97. Svirbely, J.L. Toxicity tests of decaborane for laboratory animals. I. Effect of repeated doses. AMA Arch. Ind. Hyg. Occup. Med. 1955, 11, 138–141. Available online: https://archive.org/details/sim_a-m-a-archives-of-industrial-health_1955-02_11_2/page/138/mode/2up (accessed on 31 July 2023).
  98. Walton, R.P.; Richardson, J.A.; Brodie, O.J. Cardiovascular actions of decaborane. J. Pharmacol. Exp. Ther. 1955, 114, 367–378. [Google Scholar]
  99. Lamberti, J.M. Review of the toxicological properties of pentaborane, diborane, decaborane, and boric acid. NASA Tech. Rep. 1956, NACA-RM-E56H13a. Available online: https://ntrs.nasa.gov/api/citations/19930090289/downloads/19930090289.pdf (accessed on 31 July 2023).
  100. Roush, G. The toxicology of the boranes. J. Occup. Med. 1959, 1, 46–52. Available online: https://www.jstor.org/stable/44999044 (accessed on 31 July 2023). [CrossRef]
  101. Fabre, R.; Chary, R.; Bocquet, P.; Jayot, R. Etat actuel de la toxicologie des hydrides de bore. Arch. Mal. Prof. Med. Trav. Secur. Soc. 1959, 20, 701–712. [Google Scholar]
  102. Miller, D.F.; Tamas, A.; Robinson, L.; Merriweather, E. Observations on experimental boron hydride exposures. Toxicol. Appl. Pharmacol. 1960, 2, 430–440. [Google Scholar] [CrossRef]
  103. Miller, D.F.; Tamas, A.A.; Robinson, L.; Merriweather, E. Cumulative effects of borane toxicity as revealed by a clinical test. Tech. Rep. Aerospace Med. Res. Lab. 1960, WADD-60-604. Available online: https://web.archive.org/web/20181030113902/http://www.dtic.mil/dtic/tr/fulltext/u2/247355.pdf (accessed on 31 July 2023).
  104. Delgado, J.M. Effects of decaborane on brain activity. Tech. Rep. Aerospace Med. Res. Lab. 1963, AMRL-TDR-63-41. Available online: https://web.archive.org/web/20180726163136/http://www.dtic.mil/dtic/tr/fulltext/u2/411769.pdf (accessed on 31 July 2023).
  105. Delgado, J.M.R.; Back, K.C.; Tamas, A.A. The effect of boranes on the monkey brain. Arch. Int. Pharmacodyn. Thérap. 1963, 141, 262–270. [Google Scholar]
  106. Lalli, G. Sulla tossicità di alcuni propellenti per missile. Ann. Geophys. 1963, 16, 385–406. [Google Scholar] [CrossRef]
  107. Merrit, J.A. Methylene blue in the treatment of decaborane toxicity. Arch. Environ. Health 1965, 10, 452–454. [Google Scholar] [CrossRef]
  108. Reynolds, H.H.; Back, K.C. Effect of decaborane injection on operant behavior of monkeys. Toxicol. Appl. Pharmacol. 1966, 8, 197–209. [Google Scholar] [CrossRef]
  109. Fairchild, M.D.; Sterman, M.B.; McRae, G.L. The effects of decaborane on cerebral electrical activity and locomotor behavior in the cat. Tech. Rep. Aerosp. Med. Res. Lab. 1972, AMRL-TR-72-80. Available online: https://web.archive.org/web/20180724164936/http://www.dtic.mil/dtic/tr/fulltext/u2/756526.pdf (accessed on 31 July 2023).
  110. Tadepalli, A.S.; Buckley, J.P. Cardiac and peripheral vascular effects of decaborane. Toxicol. Appl. Pharmacol. 1974, 29, 210–222. [Google Scholar] [CrossRef]
  111. Dekaboran. The MAK-Collection for Occupational Health and Safety; Greim, G., Ed.; Wiley-VCH Verlag: Berlin, Germany, 2002; Volume 1. [Google Scholar] [CrossRef]
  112. Merritt, J.H.; Schultz, E.J.; Wykes, A.A. Effect of decaborane on the norepinephrine content of rat brain. Biochem. Pharmacol. 1964, 13, 1364–1366. [Google Scholar] [CrossRef]
  113. Oliverio, A. Release of cardiac noradrenaline by decaborane in the heart-lung preparation of guinea pig. Biochem. Pharmacol. 1965, 14, 1689–1692. [Google Scholar] [CrossRef]
  114. von Euler, U.S.; Lishajko, F. Stereospecific catecholamine uptake in rabbit hearts depleted by decaborane. Int. J. Neuropharmacol. 1965, 4, 273–280. [Google Scholar] [CrossRef]
  115. von Euler, U.S.; Lishajko, F. Catecholamine depletion and uptake in adrenergic nerve vesicles and in rabbit organs after decaborane. Acta Physiol. Scand. 1965, 65, 324–330. [Google Scholar] [CrossRef]
  116. Byodeman, S.; von Euler, U.S. Neurotransmitter deficiency and reloading in noradrenaline depleted rabbits. Acta Physiol. Scand. 1966, 66, 134–140. [Google Scholar] [CrossRef]
  117. Merritt, J.H.; Schultz, E.J. The effect of decaborane on the biosynthesis and metabolism of norepinephrine in the rat brain. Life Sci. 1966, 5, 27–32. [Google Scholar] [CrossRef]
  118. Johnson, D.G. The effect of cold exposure on the catecholamine excretion of rats treated with decaborane. Acta Physiol. Scand. 1966, 68, 129–133. [Google Scholar] [CrossRef]
  119. Merritt, J.H.; Sulkowski, T.S. Inhibition of aromatic L-amino acid decarboxylation by decaborane. Biochem. Pharmacol. 1967, 16, 369–373. [Google Scholar] [CrossRef]
  120. Bhattacharya, I.C. Uptake of noradrenaline in the isolated perfused rat heart after depletion with decaborane. Acta Physiol. Scand. 1968, 73, 128–138. [Google Scholar] [CrossRef]
  121. Medina, M.A.; Landez, J.H.; Foster, L.L. Inhibition of tissue histamine formation by decaborane. J. Pharmacol. Exp. Ther. 1969, 169, 132–137. Available online: https://jpet.aspetjournals.org/content/169/1/132 (accessed on 31 July 2023).
  122. Scott, W.N.; Landez, J.H.; Cole, H.D. Effects of boranes upon tissues of the rat. I. Aspartate aminotransferase and lactic dehydrogenase. Proc. Soc. Exp. Biol. Med. 1970, 134, 348–352. [Google Scholar] [CrossRef]
  123. Korty, P.; Scott, W.N. Effects of boranes upon tissues of the rat. II. Tissue amino acid content in rats on a normal diet. Proc. Soc. Exp. Biol. Med. 1970, 135, 629–632. [Google Scholar] [CrossRef]
  124. Landez, J.H.; Scott, W.N. Effects of boranes upon tissues of the rat. III. Tissue amino acids in rats on a pyridoxine-deficient diet. Proc. Soc. Exp. Biol. Med. 1971, 136, 1389–1393. [Google Scholar] [CrossRef]
  125. Malmfors, T.; von Euler, U.S. Depletion and repletion of noradrenaline in adrenergic nerves of the rat after decaborane treatment. Experientia 1971, 27, 417–419. [Google Scholar] [CrossRef]
  126. Shahab, L.; Lishajko, F.; von Euler, U.S. Differentiated storage mechanisms for noradrenaline and dopamine in the rabbit heart. Neuropharmacology 1971, 10, 765–769. [Google Scholar] [CrossRef]
  127. Menon, M.; Clark, W.G.; Aures, D. Effect of tremorine, oxotremorine and decaborane on brain histamine levels in rats. Pharmacol. Res. Commun. 1971, 3, 345–350. [Google Scholar] [CrossRef]
  128. Schayer, R.W.; Reilly, M.A. Effect of decaborane on histamine formation in mice. J. Pharmacol. Exp. Ther. 1971, 177, 177–180. Available online: https://jpet.aspetjournals.org/content/177/1/177 (accessed on 31 July 2023).
  129. Naeger, L.L.; Leibman, K.C. Mechanisms of decaborane toxicity. Toxicol. Appl. Pharmacol. 1972, 22, 517–527. [Google Scholar] [CrossRef]
  130. Valerino, D.M.; Soliman, M.R.I.; Aurori, K.C.; Tripp, S.L.; Wykes, A.A.; Vesell, E.S. Studies on the interaction of several boron hydrides with liver microsomal enzymes. Toxicol. Appl. Pharmacol. 1974, 29, 358–366. [Google Scholar] [CrossRef]
  131. Merritt, J.A.; Meyer, H.C.; Greenberg, R.I.; Tanton, G.A. The production of decaborane-14 from diborane by laser induced chemistry. Propellants Explos. Pyrotech. 1979, 4, 78–82. [Google Scholar] [CrossRef]
  132. Dunks, G.B.; Palmer Ordonez, K. A one-step synthesis of B11H14 ion from NaBH4. Inorg. Chem. 1978, 17, 1514–1516. [Google Scholar] [CrossRef]
  133. Dunks, G.B.; Palmer Ordonez, K. A simplified preparation of B10H14 from NaBH4. Inorg. Chem. 1978, 17, 2555–2556. [Google Scholar] [CrossRef]
  134. Dunks, G.B.; Barker, K.; Hedaya, E.; Hefner, C.; Palmer-Ordonez, K.; Remec, P. Simplified synthesis of B10H14 from NaBH4 via B11H14 ion. Inorg. Chem. 1981, 20, 1692–1697. [Google Scholar] [CrossRef]
  135. Belov, P.P.; Storozhenko, P.A.; Voloshina, N.S.; Kuznetsova, M.G. Synthesis of decaborane by the reaction of sodium undecaborate with mild organic oxidants. Russ. J. Appl. Chem. 2018, 90, 1804–1809. [Google Scholar] [CrossRef]
  136. Voloshina, N.S.; Belov, P.P.; Storozhenko, P.A.; Shebashova, N.M.; Kozlova, E.E.; Egorova, N.V.; Kuznetsova, M.G.; Gurkova, E.L. Specific features of oxidation of sodium tetradecahydroundecaborate to decaborane with manganese dioxide. Russ. J. Appl. Chem. 2020, 93, 807–812. [Google Scholar] [CrossRef]
  137. Mongeot, H.; Atchekzai, H.R. Opening of the B10H102− cage to give B10H14. Z. Naturforsch. B 1981, 36, 313–314. [Google Scholar] [CrossRef]
  138. Guter, G.A.; Schaeffer, G.W. The strong acid behavior of decaborane. J. Am. Chem. Soc. 1956, 78, 3546. [Google Scholar] [CrossRef]
  139. Norment, H.J. Unit cells and space groups for two etherates of sodium tridecahydrodecaborate(1-). Acta Cryst. 1959, 12, 695. [Google Scholar] [CrossRef]
  140. Greenwood, N.N.; Sharrocks, D.N. Decaborane anions and the synthesis of polyhedral borane complexes of mercury(II) and cobalt(II). J. Chem. Soc. A 1969, 2334–2338. [Google Scholar] [CrossRef]
  141. Hawthorne, M.F.; Pitochelli, A.R.; Strahm, R.D.; Miller, J.J. The preparation and characterization of salts which contain the B10H13 anion. J. Am. Chem. Soc. 1960, 82, 1825–1829. [Google Scholar] [CrossRef]
  142. Hawthorne, M.F. The reaction of phosphine methylenes with boron hydrides. J. Am. Chem. Soc. 1958, 80, 3480–3481. [Google Scholar] [CrossRef]
  143. Onak, T.; Rosendo, H.; Siwapinyoyos, G.; Kubo, R.; Liauw, L. Reaction of 1,8-bis(dimethylamino)naphthalene, a highly basic and weakly nucleophilic amine, with several polyboranes and with boron trifluoride. Inorg. Chem. 1979, 18, 2943–2945. [Google Scholar] [CrossRef]
  144. Pérez, S.; Sanz Miguel, P.J.; Macías, R. Decaborane anion tautomerism: Ion pairing and proton transfer control. Dalton Trans. 2018, 47, 5850–5859. [Google Scholar] [CrossRef] [PubMed]
  145. Heřmánek, S.; Plotová, H.; Plešek, J. On the acidity characteristics of decaborane(14) and its benzyl derivatives in organic solvent-water systems. Collect. Czech. Chem. Commun. 1975, 40, 3593–3601. [Google Scholar] [CrossRef]
  146. McCrary, P.D.; Barber, P.S.; Kelley, S.P.; Rogers, R.D. Nonaborane and decaborane cluster anions can enhance the ignition delay in hypergolic ionic liquids and induce hypergolicity in molecular solvents. Inorg. Chem. 2014, 53, 4770–4776. [Google Scholar] [CrossRef] [PubMed]
  147. Sneddon, L.G.; Huffman, J.C.; Schaeffer, R.O.; Streib, W.E. Structure of the B10H13 ion. J. Chem. Soc. Chem. Commun. 1972, 474–475. [Google Scholar] [CrossRef]
  148. Wynd, A.J.; Welch, A.J. Structure of [PhCH2NMe3]+[B10H13]. Acta Cryst. C 1989, 45, 615–617. [Google Scholar] [CrossRef]
  149. Siedle, A.R.; Bodner, G.M.; Todd, L.J. Studies in boron hydrides—V: Assignment of the 11B NMR spectrum of the tridecahydro decaborate(1−) ion. J. Inorg. Nucl. Chem. 1971, 33, 3671–3676. [Google Scholar] [CrossRef]
  150. Wilks, P.H.; Carter, J.C. Preparation and properties of sodium decaboranate(12,2-). J. Am. Chem. Soc. 1966, 88, 3441. [Google Scholar] [CrossRef]
  151. Bridges, A.N.; Gaines, D.F. The dianion of nido-decaborane(14), nido-dodecahydrodecaborate(2-), [B10H122−], and its solution behavior. Inorg. Chem. 1995, 34, 4523–4524. [Google Scholar] [CrossRef]
  152. Hofmann, M.; von Ragué Schleyer, P. Structures of arachno- and hypho-B10 clusters and stability of their possible Lewis base adducts ([B10H12]2−, [B10H12·L]2−, [B10H12·2L]2−, [B10H13], [B10H13·L], [B10H12·2L]). An ab initio/IGLO/NMR investigation. Inorg. Chem. 1998, 37, 5557–5565. [Google Scholar] [CrossRef]
  153. Muetterties, E.L. Chemistry of boranes. VI. Preparation and structure of B10H142−. Inorg. Chem. 1963, 2, 647–648. [Google Scholar] [CrossRef]
  154. Kendall, D.S.; Lipscomb, W.N. Crystal structure of tetramethylammonium tetradecahydrodecaborate. Structure of the Bl0H142− ion. Inorg. Chem. 1973, 12, 546–551. [Google Scholar] [CrossRef]
  155. Lipscomb, W.N.; Wiersema, R.J.; Hawthorne, M.F. Structural ambiguity of the B10H142− ion. Inorg. Chem. 1972, 11, 651–652. [Google Scholar] [CrossRef]
  156. Schaeffer, R.; Tebbe, F. Formation of B10H15- as an intermediate in borohydride attack on decaborane-14. Inorg. Chem. 1964, 3, 1638–1640. [Google Scholar] [CrossRef]
  157. Rietz, R.R.; Siedle, A.R.; Schaeffer, R.O.; Todd, L.J. High-resolution nuclear magnetic resonance study of the pentadecahydro-decaborate(1-) ion. Inorg. Chem. 1973, 12, 2100–2102. [Google Scholar] [CrossRef]
  158. Zahkarkin, L.I.; Stanko, V.I.; Chapovskii, Y.A. Reactions of acetals and ortho-ethers with decaborane and diacetonitrile decaborane. Bull. Acad. Sci. USSR Div. Chem. Sci. 1962, 11, 1048–1049. [Google Scholar] [CrossRef]
  159. Lee, S.H.; Park, Y.J.; Yoon, C.M. Reductive etherification of aromatic aldehydes with decaborane. Tetrahedron Lett. 1999, 40, 6049–6050. [Google Scholar] [CrossRef]
  160. Funke, U.; Jiay, H.; Fischer, S.; Scheunemann, M.; Steinbach, J. One-step reductive etherification of 4-[18F]fluoro-benzaldehyde with decaborane. J. Label Compd. Radiopharm. 2006, 49, 745–755. [Google Scholar] [CrossRef]
  161. Park, E.S.; Lee, J.H.; Kim, S.J.; Yoon, C.M. One-pot reductive amination of acetals with aromatic amines using decaborane (B10H14) in methanol. Synth. Commun. 2003, 33, 3387–3396. [Google Scholar] [CrossRef]
  162. Toyosuke, T.; Tsuneo, M.; Katsunori, K.; Yuichi, I. The reaction of decaborane with carbonyl compounds. Bull. Chem. Soc. Jpn. 1978, 51, 1259–1260. [Google Scholar] [CrossRef]
  163. Bae, J.W.; Lee, S.H.; Jung, Y.J.; Yoon, C.-O.M.; Yoon, C.M. Reduction of ketones to alcohols using a decaborane/pyrrolidine/ cerium(III) chloride system in methanol. Tetrahedron Lett. 2001, 42, 2137–2139. [Google Scholar] [CrossRef]
  164. Lee, S.H.; Nam, M.H.; Cho, M.Y.; Yoo, B.W.; Rhee, H.J.; Yoon, C.M. Chemoselective reduction of aldehydes using decaborane in aqueous solution. Synth. Commun. 2006, 36, 2469–2474. [Google Scholar] [CrossRef]
  165. Lee, S.H.; Jung, Y.J.; Cho, Y.J.; Yoon, C.-O.M.; Hwang, H.-J.; Yoon, C.M. Dehalogenation of α-halocarbonyls using decaborane as a transfer hydrogen agent in methanol. Synth. Commun. 2001, 31, 2251–2254. [Google Scholar] [CrossRef]
  166. Lee, S.H.; Park, J.Y.; Yoon, C.M. Hydrogenation of alkenes or alkynes using decaborane in methanol. Tetrahedron Lett. 2000, 41, 887–889. [Google Scholar] [CrossRef]
  167. Hawthorne, M.F.; Miller, J.J. Deuterium exchange of decaborane with deuterium oxide and deuterium chloride. J. Am. Chem. Soc. 1958, 80, 754. [Google Scholar] [CrossRef]
  168. Miller, J.J.; Hawthorne, M.F. The course of base-catalyzed hydrogen exchange in decaborane. J. Am. Chem. Soc. 1959, 81, 4501–4503. [Google Scholar] [CrossRef]
  169. Dupont, J.A.; Hawthorne, M.F. The nature of the electrophilic deuterium exchange reaction of decaborane with deuterium chloride. J. Am. Chem. Soc. 1962, 84, 1804–1808. [Google Scholar] [CrossRef]
  170. Dupont, J.A.; Hawthorne, M.F. Deuterium exchange of decaborane with deuterium chloride under electrophilic conditions. J. Am. Chem. Soc. 1959, 81, 4998–4999. [Google Scholar] [CrossRef]
  171. Dopke, J.A.; Gaines, D.F. Deuteration of decaborane(14) via exchange with deuterated aromatic solvents. Inorg. Chem. 1999, 38, 4896–4897. [Google Scholar] [CrossRef]
  172. Gaines, D.F.; Beall, H. Hydrogen−deuterium exchange in decaborane(14):  Mechanistic studies. Inorg. Chem. 2000, 39, 1812–1813. [Google Scholar] [CrossRef] [PubMed]
  173. Hillman, M. The chemistry of decaborane. Iodination studies. J. Am. Chem. Soc. 1960, 82, 1096–1099. [Google Scholar] [CrossRef]
  174. Sprecher, R.F.; Aufderheide, B.E.; Luther, G.W., III; Carter, J.C. Boron-11 nuclear magnetic resonance chemical shift assignments for monohalogenated decaborane(14) isomers. J. Am. Chem. Soc. 1974, 96, 4404–4410. [Google Scholar] [CrossRef]
  175. Stuchlík, J. Gas chromatographic separation of 1- and 2-halogenodecaboranes. J. Chromatogr. A 1973, 81, 142–143. [Google Scholar] [CrossRef]
  176. Schaeffer, R.; Shoolery, J.N.; Jones, R. Structures of halogen substituted boranes. J. Am. Chem. Soc. 1958, 80, 2670–2673. [Google Scholar] [CrossRef]
  177. Williams, R.E.; Onak, T.P. Boron-11 nuclear magnetic resonance spectra (32.1 Mc.) of alkylated derivatives of dicarbahexaborane(8) and 1-iododecaborane(14). J. Am. Chem. Soc. 1964, 86, 3159–3160. [Google Scholar] [CrossRef]
  178. Hillman, M. The chemistry of decaborane. II. Iodination in solvent. J. Inorg. Nucl. Chem. 1960, 12, 384–385. [Google Scholar] [CrossRef]
  179. Wallbridge, M.H.G.; Williams, J.; Williams, R.L. Boron hydride derivatives. Part XI. Iodination of decaborane. J. Chem. Soc. A 1967, 132–133. [Google Scholar] [CrossRef]
  180. Safronov, A.V.; Sevryugina, Y.V.; Jalisatgi, S.S.; Kennedy, R.D.; Barnes, C.L.; Hawthorne, M.F. Unfairly forgotten member of the iodocarborane family: Synthesis and structural characterization of 8-iodo-1,2-dicarba-closo-dodecaborane, its precursors, and derivatives. Inorg. Chem. 2012, 51, 2629–2637. [Google Scholar] [CrossRef] [PubMed]
  181. Sequeira, A.; Hamilton, W.C. Crystal and molecular structure of monoiododecaborane. Inorg. Chem. 1967, 6, 1281–1286. [Google Scholar] [CrossRef]
  182. Schaeffer, R. The molecular structure of B10H12I2. J. Am. Chem. Soc. 1957, 79, 2726–2728. [Google Scholar] [CrossRef]
  183. Plešek, J.; Štíbr, B.; Heřmánek, S. Chemistry of boranes. VI. The reaction of bis-dialkylsulphido-dodecahydrodecaboranes with hydrohalogens. General preparation of 6- (or 5-) halogentridecahydrodecaboranes. Collect. Czech. Chem. Commun. 1966, 31, 4744–4745. [Google Scholar] [CrossRef]
  184. Sedmera, P.; Hanousek, F.; Samek, Z. Chemistry of boranes. XIII. Determination of structure of some halogenedecaboranes and oxido-bis-tridecahydrodecaborane by means of 11B NMR and IR spectra. Collect. Czech. Chem. Commun. 1968, 33, 2169–2175. [Google Scholar] [CrossRef]
  185. Štíbr, B.; Plešek, J.; Heřmánek, S. Chemistry of boranes. XV. Synthesis, properties, reactions and mechanism of formation of 5(6)-halogenotridecahydrodecaboranes. Collect. Czech. Chem. Commun. 1969, 34, 194–205. [Google Scholar] [CrossRef]
  186. Ewing, W.C.; Carroll, P.J.; Sneddon, L.G. Crystallographic characterizations and new high-yield synthetic routes for the complete series of 6-X-B10H13 halodecaboranes (X = F, Cl, Br, I) via superacid-induced cage-opening reactions of closo-B10H102−. Inorg. Chem. 2008, 47, 8580–8582. [Google Scholar] [CrossRef]
  187. Ewing, W.C.; Carroll, P.J.; Sneddon, L.G. Efficient Syntheses of 5-X-B10H13 halodecaboranes via the photochemical (X = I) and/or base-catalyzed (X = Cl, Br, I) isomerization reactions of 6-X-B10H13. Inorg. Chem. 2010, 49, 1983–1994. [Google Scholar] [CrossRef] [PubMed]
  188. Zakharkin, L.I.; Kalinin, V.N. Halogenation of decaborane in the presence of aluminum chloride. Zh. Obshch. Khim. 1966, 36, 2160–2162. [Google Scholar]
  189. Bonnetot, B.; Miele, P.; Naoufal, D.; Mongeot, H. The interaction of the [B10H10]2− cage with Lewis acids and the formation of decaborane derivatives by cage-opening reactions. Collect. Czech. Chem. Commun. 1997, 62, 1273–1278. [Google Scholar] [CrossRef]
  190. Stuchlík, J.; Heřmánek, S.; Plešek, J.; Štíbr, B. Chemistry of boranes. XVIII. A preparative separation of halogenodecaboranes. The isolation of 1-, 2-, 5-, and 6-bromotridecahydrodecaboranes. Collect. Czech. Chem. Commun. 1970, 35, 339–343. [Google Scholar] [CrossRef]
  191. Dupont, T.J.; Loffredo, R.E.; Haltiwanger, R.C.; Turner, C.A.; Norman, A.D. Oxidative cleavage of dimethylstanna-undecaborane: Preparation and structural characterization of 5,10-dibromodecaborane(14). Inorg. Chem. 1978, 17, 2062–2067. [Google Scholar] [CrossRef]
  192. Hillman, M.; Mangold, D.J. Chlorodecaborane. Inorg. Chem. 1965, 4, 1356–1357. [Google Scholar] [CrossRef]
  193. Williams, R.E.; Pier, E. Chlorodecaboranes identified as 1-ClB10H13 and 2-ClB10H13 by 64.2-Mc. 11B nuclear magnetic resonance spectra. Inorg. Chem. 1965, 4, 1357–1358. [Google Scholar] [CrossRef]
  194. Mongeot, H.; Atchekzai, J.; Bonnetot, B.; Colombier, M. Preparation du 6-B10H13Cl a partir de melanges AlCl3-(Et4N)2B10H10. Bull. Soc. Chim. Fr. 1987, 75–77. [Google Scholar]
  195. Bonnetot, B.; Aboukhassib, A.; Mongeot, H. Study of the interaction of AlCl3 with the B10H102− cage in the solid state. Inorg. Chim. Acta 1989, 156, 183–187. [Google Scholar] [CrossRef]
  196. Hawthorne, M.F.; Miller, J.J. The alkoxylation of decaborane. J. Am. Chem. Soc. 1960, 82, 500. [Google Scholar] [CrossRef]
  197. Norman, A.D.; Rosell, S.L. Evidence of terminal ethoxy group substitution in ethoxydecaborane(14). Inorg. Chem. 1969, 8, 2818–2820. [Google Scholar] [CrossRef]
  198. Loffredo, R.E.; Drullinger, L.F.; A. Slater, J.A.; Turner, C.A.; Norman, A.D. Preparation and properties of 6-ethoxy-, 6-phenyl-, and 6-trimethylsiloxydecaborane(14). Inorg. Chem. 1976, 15, 478–480. [Google Scholar] [CrossRef]
  199. Beachell, H.C.; Schar, W.C. The reaction of decaborane with substituted alcohols. J. Am. Chem. Soc. 1958, 80, 2943–2945. [Google Scholar] [CrossRef]
  200. Amberger, E.; Leidl, P. Synthese von (CH3)3SiB10H13. J. Organomet. Chem. 1969, 18, 345–347. [Google Scholar] [CrossRef]
  201. Ewing, W.C.; Carroll, P.J.; Sneddon, L.G. Syntheses and surprising regioselectivity of 5- and 6-substituted decaboranyl ethers via the nucleophilic attack of alcohols on 6- and 5-halodecaboranes. Inorg. Chem. 2011, 50, 4054–4064. [Google Scholar] [CrossRef]
  202. Hawthorne, M.F.; Mavunkal, I.J.; Knobler, C.B. Electrophilic reactions of protonated closo-Bl0Hl02− with arenes, alkane C-H bonds, and triflate ion forming aryl, alkyl, and triflate nido-6-X-Bl0H13 derivatives. J. Am. Chem. Soc. 1992, 114, 4427–4429. [Google Scholar] [CrossRef]
  203. Bondarev, O.; Sevryugina, Y.V.; Jalisatgi, S.S.; Hawthorne, M.F. Acid-induced opening of [closo-B10H10]2− as a new route to 6-Substituted nido-B10H13 decaboranes and related carboranes. Inorg. Chem. 2012, 51, 9935–9942. [Google Scholar] [CrossRef]
  204. Berkeley, E.R.; Ewing, W.C.; Carroll, P.J.; Sneddon, L.G. Synthesis, structural characterization, and reactivity studies of 5-CF3SO3-B10H13. Inorg. Chem. 2014, 53, 5348–5358. [Google Scholar] [CrossRef]
  205. Wang, Y.; Han, H.; Kang, J.-X.; Peng, J.; Lu, X.; Cao, H.-J.; Liu, Z.; Chen, X. Silylium ion-mediated cage-opening functionalization of closo-B10H102− salts. Chem. Commun. 2022, 58, 11933–11936. [Google Scholar] [CrossRef] [PubMed]
  206. Naoufal, D.; Kodeih, M.; Cornu, D.; Miele, P. New method of synthesis of 6-hydroxy-nido-decaborane 6-(OH)B10H13 by cage opening of closo-[B10H10]2−. J. Organomet. Chem. 2005, 690, 2787–2789. [Google Scholar] [CrossRef]
  207. Greenwood, N.N.; Hails, M.J.; Kennedy, J.D.; McDonald, W.S. Reactions of 6,6’-bis(nido-decaboranyl) oxide and 6-hydroxy-nido-decaborane with dihalogenobis(phosphine) complexes of nickel, palladium, and platinum, and some related chemistry; nuclear magnetic resonance investigations and the crystal and molecular structures of bis(dimethylphosphine)-di-µ-(2,3,4-η3-nido-hexaboranyl)-diplatinum(PtPt), [Pt2(µ-η3-B6H9)2(PMe2Ph)2], and of 2,4-dichloro-1,1-bis(dimethylphenylphosphine)-closo-1-nickeladecaborane, [(PhMe2P)2NiB9H7Cl2]. J. Chem. Soc. Dalton Trans. 1985, 953–972. [Google Scholar] [CrossRef]
  208. Štíbr, B.; Plešek, J.; Hanousek, F.; Heřmánek, S. Chemistry of boranes. XXIII. Reaction of 6,9-bis(dialkylsulfido)-dodecahydrodecaboranes with mercuric salts. Collect. Czech. Chem. Commun. 1971, 36, 1794–1799. [Google Scholar] [CrossRef]
  209. Kelley, S.P.; Rachiero, G.P.; Titi, H.M.; Rogers, R.D. New reactions for old ions: Cage rearrangements, hydrolysis, and two-electron reduction of nido-decaborane in neat 1-ethyl-3-methylimidazolium acetate. ACS Omega 2018, 3, 8491–8496. [Google Scholar] [CrossRef] [PubMed]
  210. Plešek, J.; Heřmánek, S.; Štíbr, B. Chemistry of boranes. VIII. Synthesis and reactions of 6,6’-oxido-bis-tridecahydrodecaborane. Collect. Czech. Chem. Commun. 1968, 33, 691–698. [Google Scholar] [CrossRef]
  211. Kennedy, J.D.; Greenwood, N.N. A proton and boron-l 1 NMR study of icosaborane oxide, B20H16O. Inorg. Chim. Acta 1980, 38, 93–96. [Google Scholar] [CrossRef]
  212. Greenwood, N.N.; McDonald, W.S.; Spalding, T.R. Crystal and molecular structure of 6,6’-bis(nido-decaboranyl) oxide (B10H13)2O. J. Chem. Soc. Dalton Trans. 1980, 1251–1252. [Google Scholar] [CrossRef]
  213. Bonnetot, B.; Tangi, A.; Colombier, M.; Mongeot, H. Dehydration of (H3O)2B10H10: An improved preparation of icosaborane oxide, (B10H13)2O. Inorg. Chim. Acta 1985, 105, L15–L16. [Google Scholar] [CrossRef]
  214. Pace, R.J.; Williams, J.; Williams, R.L. Boron hydride derivatives. Part VII. The characterisation of some decaborane derivatives of the type, B10H12,2M. J. Chem. Soc. 1961, 2196–2204. [Google Scholar] [CrossRef]
  215. Knoth, W.H.; Muetterties, E.L. Chemistry of boranes. II: Decaborane derivatives based on the B10H12 structural unit. J. Inorg. Nucl. Chem. 1961, 20, 66–72. [Google Scholar] [CrossRef]
  216. Fein, M.M.; Green, J.; Bobinski, J.; Cohen, M.S. Reaction products from decaborane and amides. Inorg. Chem. 1965, 4, 583–584. [Google Scholar] [CrossRef]
  217. Cragg, R.H.; Fortuin, M.S.; Greenwood, N.N. Complexes of decaborane. Part I. Ultraviolet spectra of some bis-(ligand) complexes containing phosphorus and sulphur. J. Chem. Soc. A 1970, 1817–1821. [Google Scholar] [CrossRef]
  218. Janoušek, Z.; Plešek, J.; Plzák, Z. Open cage boranes and heteroborane thiols: Syntheses, structures, and some properties. Collect. Czech. Chem. Commun. 1979, 44, 2904–2907. [Google Scholar]
  219. Bould, J.; Macháček, J.; Londesborough, M.G.S.; Macías, R.; Kennedy, J.D.; Bastl, Z.; Rupper, P.; Baše, T. Decaborane thiols as building blocks for self-assembled monolayers on metal surfaces. Inorg. Chem. 2012, 51, 1685–1694. [Google Scholar] [CrossRef]
  220. Zakharkin, L.I.; Stanko, V.I.; Okhlobystin, O.Y. Reaction of decaborane and pentaborane with mercaptans and sulfides. Bull. Acad. Sci. USSR Div. Chem. Sci. 1961, 10, 1942–1943. [Google Scholar] [CrossRef]
  221. Hawthorne, M.F.; Pilling, R.L.; Grimes, R.N. The mechanism of B10H10-2 formation from B10H12(ligand)2 species. J. Am. Chem. Soc. 1967, 89, 1067–1074. [Google Scholar] [CrossRef]
  222. Beall, H.; Gaines, D.F. Reactions of 6,9-bis(dimethyl sulfide)-decaborane(14), 6,9-[(CH3)2S]2B10H12:  Mechanistic considerations. Inorg. Chem. 1998, 37, 1420–1422. [Google Scholar] [CrossRef]
  223. Sands, D.E.; Zalkin, A. The crystal structure of B10H12[S(CH3)2]2. Acta Cryst. 1962, 15, 410–417. [Google Scholar] [CrossRef]
  224. Yumatov, V.D.; Murakhtanov, V.V.; Volkov, V.V.; Il’inchik, E.A.; Volkov, O.V. Comparative study of electronic structure of 6,9-bis(dimethyl sulfide)-arachno-decaborane(12) B10H12[S(CH3)]2 in a series of sulfide derivatives by the X-ray emission method. Russ. J. Inorg. Chem. 1998, 43, 1557–1561. [Google Scholar]
  225. Yumatov, V.D.; Il’inchik, E.A.; Mazalov, L.N.; Volkov, O.V.; Volkov, V.V. X-Ray and X-ray photoelectron spectroscopy studies of the electronic structure of borane derivatives. J. Struct. Chem. 2001, 42, 281–295. [Google Scholar] [CrossRef]
  226. Il’inchik, E.A.; Volkov, V.V.; Mazalov, L.N. X-ray photoelectron spectroscopy of boron compounds. J. Struct. Chem. 2005, 46, 523–534. [Google Scholar] [CrossRef]
  227. Volkov, V.V.; Ikorskii, V.N.; Dunaev, S.T. Diamagnetism of compounds of the B10H12L2 series. Bull. Acad. Sci. USSR Div. Chem. Sci. 1988, 37, 845–848. [Google Scholar] [CrossRef]
  228. Zakharkin, L.I.; Stanko, V.I.; Klimova, A.I. Exchange reactions of decaborane complexes of the type B10H12[X]2. Bull. Acad. Sci. USSR Div. Chem. Sci. 1964, 13, 857–858. [Google Scholar] [CrossRef]
  229. Graybill, B.M.; Hawthorne, M.F. The nature of the colored 6,9-bis-pyridine decaborane molecule, B10H12Py2. J. Am. Chem. Soc. 1961, 83, 2673–2676. [Google Scholar] [CrossRef]
  230. Marshall, M.D.; Hunt, R.M.; Hefferan, G.T.; Adams, R.M.; Makhlouf, J.M. Opening the B10H102-cage to produce B10H12(Et2S)2. J. Am. Chem. Soc. 1967, 89, 3361–3362. [Google Scholar] [CrossRef]
  231. Guillevic, G.; Dazord, J.; Mongeot, H.; Cueilleron, J. Improved conversion of potassium tetrahydroborate into bis(dialkylsulfide)-decaborane(12), B10H12(R2S)2, via bis(tetraethylammonium) decahydrodecaborate, (Et4N)2B10H10. J. Chem. Res. (S) 1978, 402. [Google Scholar]
  232. Wang, G.-C.; Lu, Y.-X.; Huang, X.-Y.; Dai, L.-X. A new method for the synthesis of bis(diethylsulfide)decaborane. Acta Chim. Sin. 1981, 39, 251–254. Available online: http://sioc-journal.cn/Jwk_hxxb/EN/Y1981/V39/I3/251 (accessed on 31 July 2023).
  233. Heying, T.L.; Naar-Colin, C. Some chemistry of substituted decaboranes. Inorg. Chem. 1964, 3, 282–285. [Google Scholar] [CrossRef]
  234. Ahmad, R.; Crook, J.E.; Greenwood, N.N.; Kennedy, J.D. Synthesis, reactions, and nuclear magnetic resonance studies of some substituted arachno-decaborane and arachno-nonaborane derivatives, and the isolation of novel polyhedral diplatinaboranes. Crystal and molecular structure of [Pt2(PMe2Ph)23-B2H5)(η3-B6H9)]. J. Chem. Soc. Dalton Trans. 1986, 2433–2442. [Google Scholar] [CrossRef]
  235. V. Petřiček, V.; Cisařova, I.; Subrtova, V. Structure of σ(+)-5-bromo-6,9-bis(dimethylsulphido)-nido-decaborane(12), C4H23B10BrS2, determined with a twinned crystal. Acta Cryst. C 1983, 39, 1070–1072. [Google Scholar] [CrossRef]
  236. Bould, J.; Dörfler, U.; Thornton-Pett, M.; Kennedy, J.D. A rearrangement of the 10-boron nido/arachno decaboranyl cluster. Inorg. Chem. Commun. 2001, 4, 544–546. [Google Scholar] [CrossRef]
  237. Bridges, A.N.; Powell, D.R.; Dopke, J.A.; Desper, J.M.; Gaines, D.F. Monoalkyldecaborane(14) syntheses via nucleophilic alkylation and hydroboration. Inorg. Chem. 1998, 37, 503–509. [Google Scholar] [CrossRef]
  238. Bould, J.; Dörfler, U.; Rath, N.P.; Barton, L.; Kilner, C.A.; Londesborough, M.G.S.; Ormsby, D.L.; Kennedy, J.D. Macropolyhedral boron-containing cluster chemistry. A synthetic approach via the auto-fusion of [6,9-(SMe2)2-arachno-B10H12]. Dalton Trans. 2006, 3752–3765. [Google Scholar] [CrossRef]
  239. Beachell, H.C.; Hoffman, D.E. The reaction of decaborane with amines and related compounds. J. Am. Chem. Soc. 1962, 84, 180–182. [Google Scholar] [CrossRef]
  240. Schaeffer, R. A new type of substituted borane. J. Am. Chem. Soc. 1957, 79, 1006–1007. [Google Scholar] [CrossRef]
  241. van der Maas Reddy, J.; Lipscomb, W.N. Molecular structure of B10H12(CH3CN)2. J. Am. Chem. Soc. 1959, 81, 754. [Google Scholar] [CrossRef]
  242. van der Maas Reddy, J.; Lipscomb, W.N. Molecular structure of B10H12(CH3CN)2. J. Chem. Phys. 1959, 31, 610–616. [Google Scholar] [CrossRef]
  243. Mebs, S.; Kalinowski, R.; Grabowsky, S.; Förster, D.; Kickbusch, R.; Justus, E.; Morgenroth, W.; Paulmann, C.; Luger, P.; Gabel, D.; et al. Real-space indicators for chemical bonding. Experimental and theoretical electron density studies of four deltahedral boranes. Inorg. Chem. 2011, 50, 90–103. [Google Scholar] [CrossRef] [PubMed]
  244. Beall, H. Icosahedral carboranes. XVII. Simplified preparation of o-carborane. Inorg. Chem. 1972, 11, 637–638. [Google Scholar] [CrossRef]
  245. Hawthorne, M.F.; Pitochelli, A.R. Displacement reactions on the B10H12 unit. J. Am. Chem. Soc. 1958, 80, 6685. [Google Scholar] [CrossRef]
  246. Hawthorne, M.F.; Pitochelli, A.R. The reactions of bis-acetonitrile decaborane with amines. J. Am. Chem. Soc. 1959, 81, 5519. [Google Scholar] [CrossRef]
  247. Fein, M.M.; Bobinski, J.; Paustian, J.E.; Grafstein, D.; Cohen, M.S. Reaction of decaborane and its derivatives. II. Addition reactions of 6,9-bis(acetonitrile)decaborane with hydrazine. Inorg. Chem. 1965, 4, 422. [Google Scholar] [CrossRef]
  248. Froehner, G.; Challis, K.; Gagnon, K.; Getman, T.D.; Luck, R.L. A re-investigation of the reactions of amines and alcohols with 6,9-bis-(acetonitrile)decaborane. Synth. React. Inorg. Metal-Org. Nano-Metal Chem. 2007, 36, 777–785. [Google Scholar] [CrossRef]
  249. Stogniy, M.Y.; Erokhina, S.A.; Sivaev, I.B.; Bregadze, V.I. Nitrilium derivatives of polyhedral boron compounds (boranes, carboranes, metallacarboranes): Synthesis and reactivity. Phosphorus Sulfur Silicon Relat. Elem. 2019, 194, 983–988. [Google Scholar] [CrossRef]
  250. Zhdanov, A.P.; Nelyubin, A.V.; Klyukin, I.N.; Selivanov, N.A.; Bortnikov, E.O.; Grigoriev, M.S.; Zhizhin, K.Y.; Kuznetsov, N.T. Nucleophilic addition reaction of secondary amines to acetonitrilium closo-decaborate [2-B10H9NCCH3]. Russ. J. Inorg. Chem. 2019, 64, 841–846. [Google Scholar] [CrossRef]
  251. Stogniy, M.Y.; Erokhina, S.A.; Suponitsky, K.Y.; Anisimov, A.A.; Godovikov, I.A.; Sivaev, I.B.; Bregadze, V.I. Synthesis of novel carboranyl amidines. J. Organomet. Chem. 2020, 909, 121111. [Google Scholar] [CrossRef]
  252. Bogdanova, E.V.; Stogniy, M.Y.; Chekulaeva, L.A.; Anisimov, A.A.; Suponitsky, K.Y.; Sivaev, I.B.; Grin, M.A.; Mironov, A.F.; Bregadze, V.I. Synthesis and reactivity of propionitrilium derivatives of cobalt and iron bis(dicarbollides). New J. Chem. 2020, 44, 15836–15848. [Google Scholar] [CrossRef]
  253. El Anwar, S.; Růžičková, Z.; Bavol, D.; Fojt, L.; Grüner, B. Tetrazole ring substitution at carbon and boron sites of the cobalt bis(dicarbollide) ion available via dipolar cycloadditions. Inorg. Chem. 2020, 59, 17430–17442. [Google Scholar] [CrossRef] [PubMed]
  254. Bogdanova, E.V.; Stogniy, M.Y.; Suponitsky, K.Y.; Sivaev, I.B.; Bregadze, V.I. Synthesis of boronated amidines by addition of amines to nitrilium derivative of cobalt bis(dicarbollide). Molecules 2021, 26, 6544. [Google Scholar] [CrossRef] [PubMed]
  255. Voinova, V.V.; Selivanov, N.A.; Plyushchenko, I.V.; Vokuev, M.F.; Bykov, A.Y.; Klyukin, I.N.; Novikov, A.S.; Zhdanov, A.P.; Grigoriev, M.S.; Rodin, I.A.; et al. Fused 1,2-diboraoxazoles based on closo-decaborate anion—Novel members of diboroheterocycle class. Molecules 2021, 26, 248. [Google Scholar] [CrossRef] [PubMed]
  256. Stogniy, M.Y.; Erokhina, S.A.; Suponitsky, K.Y.; Markov, V.Y.; Sivaev, I.B. Synthesis and crystal structures of nickel(II) and palladium(II) complexes with o-carboranyl amidine ligands. Dalton Trans. 2021, 50, 4967–4975. [Google Scholar] [CrossRef]
  257. Nelyubin, A.V.; Selivanov, N.A.; Bykov, A.Y.; Klyukin, I.N.; Novikov, A.S.; Zhdanov, A.P.; Karpechenko, N.Y.; Grigoriev, M.S.; Zhizhin, K.Y.; Kuznetsov, N.T. Primary amine nucleophilic addition to nitrilium closo-dodecaborate [B12H11NCCH3]: A simple and effective route to the new BNCT drug design. Int. J. Mol. Sci. 2021, 22, 13391. [Google Scholar] [CrossRef]
  258. Laskova, J.; Ananiev, I.; Kosenko, I.; Serdyukov, A.; Stogniy, M.; Sivaev, I.; Grin, M.; Semioshkin, A.; Bregadze, V.I. Nucleophilic addition reactions to nitrilium derivatives [B12H11NCCH3] and [B12H11NCCH2CH3]. Synthesis and structures of closo-dodecaborate-based iminols, amides and amidines. Dalton Trans. 2022, 51, 3051–3059. [Google Scholar] [CrossRef]
  259. Williams, J.; Williams, R.L.; Wright, J.C. Boron hydride derivatives. Part IX. The reaction of decaborane with ammonia. J. Chem. Soc. 1963, 5816–5824. [Google Scholar] [CrossRef]
  260. Rozenberg, A.S.; Neehiporenko, G.N.; Alekseev, A.P. Decomposition of nitrogen-hydrogen complexes of decaborane(12). 1. The kinetics of the thermal decomposition of decaborane(12) diammoniate. Bull. Acad. Sci. USSR Div. Chem. Sci. 1978, 27, 284–287. [Google Scholar] [CrossRef]
  261. Baidina, I.A.; Podberezskaya, N.V.; Alekseev, V.I.; Volkov, V.V.; Borisov, S.V. Crystal structure of 6,9-bis-aminododecahydro-nido-decaborane B10H12(NH3)2. J. Struct. Chem. 1978, 19, 476–479. [Google Scholar] [CrossRef]
  262. Polyanskaya, T.M.; Volkov, V.V. Crystal and molecular structure of 6,9-bis(trimethylamino)-nido-decaborane(12) B10H12[N(CH3)3]2. J. Struct. Chem. 1989, 30, 629–634. [Google Scholar] [CrossRef]
  263. Baidina, I.A.; Podberezskaya, N.V.; Volkov, V.V.; Borisov, S.V. Crystal structure of 6,9-bis-(triethylamino)-nido-decaborane B10H12[(C2H5)3N]2. J. Struct. Chem. 1978, 19, 479–482. [Google Scholar] [CrossRef]
  264. Volkov, V.V.; Il’inchik, E.A.; Khudorozhko, G.F.; Yumatov, V.D.; Mazalov, L.N. Comparative study of B10H12(NH3)2 and B10H12(NEt3)2. Bull. Acad. Sci. USSR Div. Chem. Sci. 1985, 34, 2079–2085. [Google Scholar] [CrossRef]
  265. Rozenberg, A.S. Decomposition of nitrogen-hydrogen complexes of decaborane(12). 2. An IR study of the mechanism of decomposition of decaborane(12) diammoniate. Bull. Acad. Sci. USSR Div. Chem. Sci. 1978, 27, 287–290. [Google Scholar] [CrossRef]
  266. Isaenko, L.I.; Myakishev, K.G.; Posnaya, I.S.; Volkov, V.V. About the thermal stability of several derivatives of boron hydrides. Izv. Sib. Otd. Akad. Nauk SSSR Ser. Khim. 1982, 73–78. [Google Scholar]
  267. Graybill, B.M.; Pitochelli, A.R.; Hawthorne, M.F. The preparation and reactions of B10H13(Ligand) anions. Inorg. Chem. 1962, 1, 622–626. [Google Scholar] [CrossRef]
  268. Cendrowski-Guillaume, S.M.; O’Loughlin, J.L.; Pelczer, I.; Spencer, J.T. Reactivity of decaborane(14) with pyridine: Synthesis and characterization of the first 6,6-substituted isomer of nido-B10H14, 6,6-(C5H5N)2B10H12, and application of 11B-11B double-quantum NMR spectroscopy. Inorg. Chem. 1995, 34, 3935–3941. [Google Scholar] [CrossRef]
  269. Bould, J.; Laromaine, A.; Bullen, N.J.; Viñas, C.; Thornton-Pett, M.; Sillanpää, R.; Kivekäs, R.; Kennedy, J.D.; Teixidor, F. Borane reaction chemistry. Alkyne insertion reactions into boron-containing clusters. Products from the thermolysis of [6,9-(2-HC≡C–C5H4N)2-arachno-B10H12]. Dalton Trans. 2008, 1552–1563. [Google Scholar] [CrossRef]
  270. Volkov, V.V.; Myakishev, K.G.; Potapova, O.G.; Polyanskaya, T.M.; Dunaev, S.T.; Il’inchik, E.A.; Baidina, I.A.; Hudorozhko, G.F. Synthesis and physico-chemical study of 6,9-bis-pyridino-nido-decaborane(14). Izv. Sib. Otd. Akad. Nauk SSSR Ser. Khim. 1988, 3, 20–27. [Google Scholar]
  271. Il’inchik, E.A.; Volkov, V.V.; Dunaev, S.T. Structural effects in the electron absorption and luminescence spectra of decaborane(14) derivatives B10H12L2. J. Struct. Chem. 1996, 37, 51–57. [Google Scholar] [CrossRef]
  272. Volkov, V.V.; Il’inchik, E.A.; Volkov, O.V.; Yuryeva, O.P. The luminescence of cluster derivatives of boron hydrides, and some applied aspects. Chem. Sustain. Develop. 2000, 8, 185–191. Available online: https://sibran.ru/upload/iblock/fed/the_luminescence_of_cluster_derivatives_of_boron_hydrides_and_some_applied_aspects.PDF (accessed on 31 July 2023).
  273. Volkov, V.V.; Il’inchik, E.A.; Yur’eva, O.P.; Volkov, O.V. Luminescence of the derivatives of boranes and adducts of decaborane(14) of the B10H12[Py(X)]2 type. J. Appl. Spectr. 2000, 67, 864–870. [Google Scholar] [CrossRef]
  274. Volkov, V.V.; Myakishev, K.G.; Dunaev, S.T. Thermal transformations of B10H12(NH3)2, B10H12(C5H5N)2, and B10H12(C9H7N)2. Bull. Acad. Sci. USSR Div. Chem. Sci. 1988, 37, 2234–2236. [Google Scholar] [CrossRef]
  275. Polyanskaya, T.M.; Volkov, V.V.; Andrianov, V.I.; Il’inchik, E.A. Structures of two modifications of 6,9-bis-pyridine-nido-decaborane(12). Bull. Acad. Sci. USSR Div. Chem. Sci. 1989, 38, 1751–1754. [Google Scholar] [CrossRef]
  276. Londesborough, M.G.S.; Price, C.; Thornton-Pett, M.; Clegg, W.; Kennedy, J.D. Two potential pyridine-borane oligomer and polymer building blocks. Structural characterisation of [NC5H4·C5H4N·B10H12·NC5H4·C5H4N] and [Me2S·B10H12·NC4H4N·B10H12·SMe2] by conventional and synchrotron X-ray methods. Inorg. Chem. Commun. 1999, 2, 298–300. [Google Scholar] [CrossRef]
  277. Genady, A.R.; Fayed, T.A.; Gabel, D. Synthesis, characterization, and spectrophotometric studies of novel fluorescent arachno decaborane and nonaborane clusters containing aza-distyrylbenzene derivatives. J. Organomet. Chem. 2008, 693, 1065–1072. [Google Scholar] [CrossRef]
  278. Hawthorne, M.F.; Pilling, R.L.; Vasavada, R.C. The mechanism of ligand exchange with B10H12(ligand)2 species. J. Am. Chem. Soc. 1967, 89, 1075–1078. [Google Scholar] [CrossRef]
  279. Rachiero, G.P.; Titi, H.M.; Rogers, R.D. Versatility and remarkable hypergolicity of exo-6, exo-9 imidazole-substituted nido-decaborane. Chem. Commun. 2017, 53, 7736–7739. [Google Scholar] [CrossRef]
  280. Fetter, N.R. Reaction of decaborane with 2-isopropyl- and 2-methyl-5-(2-chloroethyl) tetrazole. Chem. Ind. 1959, 1548. [Google Scholar]
  281. Kendall, D.S.; Lipscomb, W.N. Molecular structure and two crystal structures of 6-isothiocyanodecaborane, 6-B10H13NCS. Inorg. Chem. 1973, 12, 2915–2919. [Google Scholar] [CrossRef]
  282. Müller, J.; Paetzold, P.; Boese, R. The reaction of decaborane with hydrazoic acid: A novel access to azaboranes. Heteroat. Chem. 1990, 1, 461–465. [Google Scholar] [CrossRef]
  283. Il’inchik, E.A.; Dunaev, S.T.; Myakishev, K.G.; Asanov, I.P. On solvatation and thermal transformations of B10H12(PPh3)2. Zh. Neorg. Khim. 1994, 39, 1071–1074. [Google Scholar]
  284. Polyanskaya, T.M.; Yumatov, V.D.; Volkov, V.V. Molecular and electronic structure of 6,9-(PPh3)2-arachno-B10H12. Dokl. Chem. 2003, 390, 144–147. [Google Scholar] [CrossRef]
  285. Fontaine, X.L.R.; Kennedy, J.D. Identification of the endo,exo isomer of 6,9-(PMe2Ph)2-arachno-B10H12 by nuclear magnetic resonance spectroscopy. J. Chem. Soc. Dalton Trans. 1987, 1573–1575. [Google Scholar] [CrossRef]
  286. Dörfler, U.; McGrath, T.D.; Cooke, P.A.; Kennedy, J.D.; Thornton-Pett, M. exo,endo and exo,exo isomers of 6,9-(PMe2Ph)2-arachno-B10H12 and its halogenated derivatives. Molecular structures of exo,endo- and exo,exo-6,9-(PMe2Ph)2-arachno-B10H12 and exo-6,endo-9-(PMe2Ph)2-2-Br-arachno-B10H11. J. Chem. Soc. Dalton Trans. 1997, 4739–4746. [Google Scholar] [CrossRef]
  287. Zakharkin, L.I.; Stanko, V.I. Complexes of decaborane with organic phosphorus and arsenic compounds. Bull. Acad. Sci. USSR Div. Chem. Sci. 1961, 10, 1936–1937. [Google Scholar] [CrossRef]
  288. Polak, R.J.; Heying, T.L. The preparation of phosphite and phosphinite decaboranes. J. Org. Chem. 1962, 27, 1483–1484. [Google Scholar] [CrossRef]
  289. Stanko, V.I.; Klimova, A.I.; Zakharkin, L.I. Complexes of decaborane with trialkyl-, triaryl-, trialkyltrithiophosphites and trialkyl-, trialkyltrithioarsenites. Bull. Acad. Sci. USSR Div. Chem. Sci. 1962, 11, 856–857. [Google Scholar] [CrossRef]
  290. Schroeder, H.; Reiner, J.R.; Heying, T.L. Chemistry of decaborane-phosphorus compounds. I. Nucleophilic substitutions of bis-(chlorodiphenylphosphine)-decaborane. Inorg. Chem. 1962, 1, 618–621. [Google Scholar] [CrossRef]
  291. Schroeder, H.; Reiner, J.R.; Knowles, T.A. Chemistry of decaborane-phosphorus compounds. III. Decaborane-14-phosphine polymers. Inorg. Chem. 1963, 2, 393–396. [Google Scholar] [CrossRef]
  292. Seyferth, D.; Rees, W.S.; Haggerty, J.S.; Lightfoot, A. Preparation of boron-containing ceramic materials by pyrolysis of the decaborane(14)-derived [B10H12·Ph2POPPh2]x polymer. Chem. Mater. 1989, 1, 45–52. [Google Scholar] [CrossRef]
  293. Packirisamy, S. Decaborane(14)-based polymers. Progr. Polym. Sci. 1996, 21, 707–773. [Google Scholar] [CrossRef]
  294. Donaghy, K.J.; Carroll, P.J.; Sneddon, L.G. Reactions of 1,1’-bis(diphenylphosphino)ferrocene with boranes, thiaboranes, and carboranes. Inorg. Chem. 1997, 36, 547–553. [Google Scholar] [CrossRef]
  295. Schroeder, H. Chemistry of decaborane-phosphorus compounds. II. Synthesis and reactions of diphenylphosphinodecaborane-14. Inorg. Chem. 1963, 2, 390–393. [Google Scholar] [CrossRef]
  296. Friedman, L.B.; Perry, S.L. Crystal and molecular structure of 5,6-μ-diphenylphosphino-decaborane(14). Inorg. Chem. 1973, 12, 288–293. [Google Scholar] [CrossRef]
  297. Muetterties, E.L.; Aftandilian, V.D. Chemistry of boranes. IV. Phosphine derivatives of B10H14 and B9H15. Inorg. Chem. 1962, 1, 731–734. [Google Scholar] [CrossRef]
  298. Thornton-Pett, M.; Beckett, M.A.; Kennedy, J.D. Polyhedral phosphaborane chemistry: Crystal and molecular structure of the diphenylphosphido-bridged arachno-decaboranyl cluster compound [PMePh3][6,9-µ-(PPh2)B10H12]. J. Chem. Soc. Dalton Trans. 1986, 303–308. [Google Scholar] [CrossRef]
  299. Miller, R.W.; Spencer, J.T. Small heteroborane cluster systems. 7. Reaction of phosphaalkyne t-BuCP with bis-(acetonitrile)decaborane(12). A new synthetic route to a large phosphaborane cluster compound. Polyhedron 1996, 15, 3151–3155. [Google Scholar] [CrossRef]
  300. Miller, R.W.; Spencer, J.T. Small heteroborane cluster systems. 8. Preparation of phosphaborane clusters from the reaction of polyhedral boranes with low-coordinate phosphorus compounds:  Reaction chemistry of phosphaalkynes with decaborane(14). Organometallics 1996, 15, 4293–4300. [Google Scholar] [CrossRef]
  301. Williams, R.L.; Dunstan, I.; Blay, N.J. Boron hydride derivatives. Part IV. Friedel–Crafts methylation of decaborane. J. Chem. Soc. 1960, 5006. [Google Scholar] [CrossRef]
  302. Obenland, C.O.; Newberry, J.R.; Schreiner, W.L.; Bartoszek, E.J. Friedel-Crafts methylation of decaborane. Ind. Eng. Chem. Prod. Res. Dev. 1965, 4, 281–283. [Google Scholar] [CrossRef]
  303. Polak, R.J.; Goodspeed, N.C. Catalyst study in methylation of decaborane. Ind. Eng. Chem. Prod. Res. Dev. 1965, 4, 158–160. [Google Scholar] [CrossRef]
  304. Holub, J.; Růžička, A.; Růžičková, Z.; Fanfrlík, J.; Hnyk, D.; Štíbr, B. Electrophilic methylation of decaborane(14): Selective synthesis of tetramethylated and heptamethylated decaboranes and their conjugated bases. Inorg. Chem. 2020, 59, 10540–10547. [Google Scholar] [CrossRef] [PubMed]
  305. Blay, N.J.; Dunstan, I.; Williams, R.L. Boron hydride derivatives. Part III. Electrophilic substitution in pentaborane and decaborane. J. Chem. Soc. 1960, 430–433. [Google Scholar] [CrossRef]
  306. Cueilleron, J.; Guillot, P. Preparation de quelques composes du decaborane. Bull. Chim. Fr. 1960, 2044–2052. [Google Scholar]
  307. Perloff, A. The crystal structure of 1-ethyldecaborane. Acta Cryst. 1964, 17, 332–338. [Google Scholar] [CrossRef]
  308. Dunstan, I.; Williams, R.L.; Blay, N.J. Boron hydride derivatives. Part V. Nucleophilic substitution in decaborane. J. Chem. Soc. 1960, 5012–5015. [Google Scholar] [CrossRef]
  309. Dunstan, I.; Blay, N.J.; Williams, R.L. Boron hydride derivatives. Part VI. Decaborane Grignard reagent. J. Chem. Soc. 1960, 5016–5019. [Google Scholar] [CrossRef]
  310. Gallaghan, J.; Siegel, B. Grignard synthesis of alkyl decaboranes. J. Am. Chem. Soc. 1959, 81, 504. [Google Scholar] [CrossRef]
  311. Siegel, B.; Mack, J.L.; Lowe, J.U.; Gallaghan, J. Decaborane Grignard reagents. J. Am. Chem. Soc. 1958, 80, 4523–4526. [Google Scholar] [CrossRef]
  312. Palchak, R.J.F.; Norman, J.H.; Williams, R.E. Decaborane, “6-benzyl” B10H13 chemistry. J. Am. Chem. Soc. 1961, 83, 3380–3384. [Google Scholar] [CrossRef]
  313. Gaines, D.F.; Bridges, A.N. New routes to monoalkyl decaborane(14) derivatives. Organometallics 1993, 12, 2015–2016. [Google Scholar] [CrossRef]
  314. Tolpin, E.I.; Mizusawa, E.; Becker, D.S.; Venzel, J. Synthesis and chemistry of 9-cyclohexyl-5(7)-(dimethyl sulfide)-nido-decaborane(11), B10H11C6H11S(CH3)2. Inorg. Chem. 1980, 19, 1182–1187. [Google Scholar] [CrossRef]
  315. Millan, M.D.; Davis, J.H. Hydroboration of (1R)-(+)-α-pinene and (1S)-(−)-β-pinene with B10H12(SMe2)2: A straightforward approach to the preparation of optically active 6-(alkyl)-nido-B10H13 derivatives. Tetrahedron Asymmetry 1998, 9, 709–712. [Google Scholar] [CrossRef]
  316. Mizusawa, E.; Rudnick, S.E.; Eriks, K. The crystal and molecular structure of 9-cyclohexyl-5(7)-(dimethyl sulfide)-nido-decaborane(11), B10H11C6H11S(CH3)2. Inorg. Chem. 1980, 19, 1188–1191. [Google Scholar] [CrossRef]
  317. Kusari, U.; Li, Y.; Bradley, M.G.; Sneddon, L.G. Polyborane reactions in ionic liquids:  New efficient routes to functionalized decaborane and o-carborane clusters. J. Am. Chem. Soc. 2004, 126, 8662–8663. [Google Scholar] [CrossRef] [PubMed]
  318. Kusari, U.; Carroll, P.J.; Sneddon, L.G. Ionic-liquid-promoted decaborane olefin-hydroboration: A new efficient route to 6-R-B10H13 derivatives. Inorg. Chem. 2008, 47, 9203–9215. [Google Scholar] [CrossRef]
  319. Yu, X.-H.; Cao, K.; Huang, Y.; Yang, J.; Li, J.; Chang, G. Platinum catalyzed sequential hydroboration of decaborane: A facile approach to poly(alkenyldecaborane) with decaborane in the mainchain. Chem. Commun. 2014, 50, 4585–4587. [Google Scholar] [CrossRef]
  320. Boggio, P.; Toppino, A.; Geninatti Crich, S.; Alberti, D.; Marabello, D.; Medana, C.; Prandi, C.; Venturello, P.; Aime, S.; Deagostino, A. The hydroboration reaction as a key for a straightforward synthesis of new MRI-NCT agents. Org. Biomol. Chem. 2015, 13, 3288–3297. [Google Scholar] [CrossRef]
  321. Naoufal, D.; Laila, Z.; Yazbeck, O.; Hamad, H.; Ibrahim, G.; Aoun, R.; Safa, A.; El Jamal, M. Kanj Synthesis, characterization and mechanism of formation of 6-substituted nido-B10H13 decaboranes by the opening reaction of closo-decahydrodecaborate [B10H10]2- cage. Main Group Chem. 2013, 12, 39–48. [Google Scholar] [CrossRef]
  322. Pender, M.J.; Wideman, T.; Carroll, P.J.; Sneddon, L.G. Transition metal promoted reactions of boron hydrides. 15. Titanium-catalyzed decaborane−olefin hydroborations:  One-step, high-yield syntheses of monoalkyldecaboranes. J. Am. Chem. Soc. 1998, 120, 9108–9109. [Google Scholar] [CrossRef]
  323. Pender, M.J.; Carroll, P.J.; Sneddon, L.G. Transition-metal-promoted reactions of boron hydrides. 17. Titanium-catalyzed decaborane−olefin hydroborations. J. Am. Chem. Soc. 2001, 123, 12222–12231. [Google Scholar] [CrossRef] [PubMed]
  324. Wei, X.; Carroll, P.J.; Sneddon, L.G. New routes to organodecaborane polymers via ruthenium-catalyzed ring-opening metathesis polymerization. Organometallics 2004, 23, 163–165. [Google Scholar] [CrossRef]
  325. Wei, X.; Carroll, P.J.; Sneddon, L.G. Ruthenium-catalyzed ring-opening polymerization syntheses of poly(organodecaboranes):  New single-source boron-carbide precursors. Chem. Mater. 2006, 18, 1113–1123. [Google Scholar] [CrossRef]
  326. Pender, M.J.; Forsthoefel, K.M.; Sneddon, L.G. Molecular and polymeric precursors to boron carbide nanofibers, nanocylinders, and nanoporous ceramics. Pure Appl. Chem. 2003, 75, 1287–1294. [Google Scholar] [CrossRef]
  327. Pender, M.J.; Sneddon, L.G. An efficient template synthesis of aligned boron carbide nanofibers using a single-source molecular precursor. Chem. Mater. 2000, 12, 280–283. [Google Scholar] [CrossRef]
  328. Li, J.; Yang, H.; Liang, Y.; Chen, J.; Xie, M.; Zhou, L.; Xiong, Q.; Li, W. Ruthenium-catalyzed cascade ring-opening polymerization/hydroboration for the synthesis of low cross-linking poly(6-norbornenyldecaborane) and its thermal property. Adv. Mater. Res. 2014, 1035, 288–291. [Google Scholar]
  329. Zhang, X.; Li, J.; Cao, K.; Yi, Y.; Yang, J.; Li, B. Synthesis and characterization of B-C polymer hollow microspheres from a new organodecaborane preceramic polymer. RSC Adv. 2015, 5, 86214–86218. [Google Scholar] [CrossRef]
  330. Wang, J.; Gou, Y.; Jian, K.; Huang, J.; Wang, H. Boron carbide ceramic hollow microspheres prepared from poly(6-CH2=CH(CH2)4-B10H13) precursor. Mater. Design 2016, 109, 408–414. [Google Scholar] [CrossRef]
  331. Wang, J.; Gou, Y.; Zhang, Q.; Jian, K.; Chen, Z.; Wang, H. Linear organodecaborane block copolymer as a single-source precursor for porous boron carbide ceramics. J. Eur. Ceram. Soc. 2017, 37, 1937–1943. [Google Scholar] [CrossRef]
  332. Li, J.; Cao, K.; Li, J.; Liu, M.; Zhang, S.; Yang, J.; Zhang, Z.; Li, B. Synthesis and ceramic conversion of a new organodecaborane preceramic polymer with high-ceramic-yield. Molecules 2018, 23, 2461. [Google Scholar] [CrossRef]
  333. Mazighi, K.; Carroll, P.J.; Sneddon, L.G. Transition metal promoted reactions of boron hydrides. 13. Platinum catalyzed synthesis of 6,9-dialkyldecaboranes. Inorg. Chem. 1993, 32, 1963–1969. [Google Scholar] [CrossRef]
  334. Chatterjee, S.; Carroll, P.J.; Sneddon, L.G. Metal-catalyzed decaborane-alkyne hydroboration reactions: Efficient routes to alkenyldecaboranes. Inorg. Chem. 2010, 49, 3095–3097. [Google Scholar] [CrossRef] [PubMed]
  335. Chatterjee, S.; Carroll, P.J.; Sneddon, L.G. Iridium and ruthenium catalyzed syntheses, hydroborations, and metathesis reactions of alkenyl-decaboranes. Inorg. Chem. 2013, 52, 9119–9130. [Google Scholar] [CrossRef] [PubMed]
  336. Ernest, R.L.; Quintana, W.; Rosen, R.; Carroll, P.J.; Sneddon, L.G. Reactions of decaborane(14) with silylated acetylenes. Synthesis of the new monocarbon carborane 9-Me2S-7-[(Me3Si)2CH]CB10H11. Organometallics 1987, 6, 80–88. [Google Scholar] [CrossRef]
  337. Burgos-Adorno, G.; Carroll, P.J.; Quintana, W. Synthesis and characterization of a new alkenyldecaborane and alkenyl monocarbon carboranes. Inorg. Chem. 1996, 35, 2568–2575. [Google Scholar] [CrossRef]
  338. Meyer, F.; Paetzold, P.; Englert, U. Reaktion von Decaboran mit dem Phosphaalkin PCtBu. Chem. Ber. 1994, 127, 93–95. [Google Scholar] [CrossRef]
  339. Bould, J.; Londesborough, M.G.S.; Ormsby, D.L.; MacBride, J.A.H.; Wade, K.; Kilner, C.A.; Clegg, W.; Teat, S.J.; Thornton-Pett, M.; Greatrex, R.; et al. Macropolyhedral boron-containing cluster chemistry: Models for intermediates en route to globular and discoidal megaloborane assemblies. Structures of [nido-B10H12(nido-B5H8)2] and [(CH2CH2C5H4N)-arachno-B10H10(NC5H4-closo-C2B10H10)] as determined by synchrotron X-ray diffraction analysis. J. Organomet. Chem. 2002, 657, 256–261. [Google Scholar] [CrossRef]
  340. Kim, S.; Treacy, J.W.; Nelson, Y.A.; Gonzalez, J.A.M.; Gembicky, M.; Houk, K.N.; Spokoyny, A.M. Arene C–H borylation strategy enabled by a non-classical boron cluster-based electrophile. Nat. Commun. 2023, 14, 1671. [Google Scholar] [CrossRef]
  341. Demel, J.; Kloda, M.; Lang, K.; Škoch, K.; Hynek, J.; Opravil, A.; Novotný, M.; Bould, J.; Ehn, M.; Londesborough, M.G.S. Direct Phenylation of nido-B10H14. J. Org. Chem. 2022, 87, 10034–10043. [Google Scholar] [CrossRef]
  342. Bůžek, D.; Škoch, K.; Ondrušová, S.; Kloda, M.; Bavol, D.; Mahun, A.; Kobera, L.; Lang, K.; Londesborough, M.G.S.; Demel, J. “Activated Borane”—A porous borane cluster network as an effective adsorbent for removing organic pollutants. Chem. Eur. J. 2022, 28, e202201885. [Google Scholar] [CrossRef]
  343. Wille, A.E.; Su, K.; Carroll, P.J.; Sneddon, L.G. New synthetic routes to azacarborane clusters:  Nitrile insertion reactions of nido-5,6-C2B8H11- and nido-B10H13. J. Am. Chem. Soc. 1996, 118, 6407–6421. [Google Scholar] [CrossRef]
  344. Baše, K.; Alcock, N.W.; Howarth, O.W.; Powell, H.R.; Harrison, A.T.; Wallbridge, M.G.H. The structure of the arachno-[B10H13C≡N]2− anion: An example of endo substitution in the decaborane(l4) framework. J. Chem. Soc. Chem. Commun. 1988, 341–342. [Google Scholar] [CrossRef]
  345. Orlova, A.M.; Sivaev, I.B.; Lagun, V.L.; Katser, S.B.; Solntsev, K.A.; Kuznetsov, N.T. Synthesis and structure of Pb(Bipy)2(6-B10H13CN). Koord. Khim. 1996, 22, 119–124. [Google Scholar]
  346. Sivaev, I.B. Chemistry of 11-vertex polyhedral boron hydrides (Review). Russ. J. Inorg. Chem. 2019, 64, 955–976. [Google Scholar] [CrossRef]
  347. Geisberger, G.; Linti, G.; Nöth, H. Beiträge zur Chemie des Bors. 213. Reaktionen eines Amino-imino-borans mit Triboran(7) und Decaboran(14). Chem. Ber. 1992, 125, 2691–2699. [Google Scholar] [CrossRef]
  348. Bridges, A.N.; Liu, J.; Kultyshev, R.G.; Gaines, D.F.; Shore, S.G. Partial insertion of the 9-BBN unit into the nido-B10 framework:  Preparation and structural characterization of (9-BBN)B10H13 and [(9-BBN)B10H12]. Inorg. Chem. 1998, 37, 3276–3283. [Google Scholar] [CrossRef]
  349. Greenwood, N.N.; Kennedy, J.D.; Taylorson, D. Mass spectroscopic evidence for icosaborane(26). J. Phys. Chem. 1978, 82, 623–625. [Google Scholar] [CrossRef]
  350. Greenwood, N.N.; Kennedy, J.D.; Spalding, T.R.; Taylorson, D. Isomers of icosaborane(26): Some synthetic routes and preliminary characterisations in the bis(nido-decaboranyl) system. J. Chem. Soc. Dalton Trans. 1979, 840–846. [Google Scholar] [CrossRef]
  351. Boocock, S.K.; Cheek, Y.M.; Greenwood, N.N.; Kennedy, J.D. A new route to isomers of icosaborane(26), B20H26. The use of 115.5-MHz 11B and 11B-{1H} nuclear magnetic resonance spectroscopy for the comparison and characterisation of separated isomers and the identification of three further icosaboranes as 1,2’-, 2,5’-, and 5,5’(or 5,7’)-(B10H13)2. J. Chem. Soc. Dalton Trans. 1981, 1430–1437. [Google Scholar] [CrossRef]
  352. Brown, G.M.; Pinson, J.W.; Ingram, L.L. Crystal and molecular structure of 1,5’-bidecaboran(14)yl: A new borane from γ irradiation of decaborane(14). Inorg. Chem. 1979, 18, 1951–1956. [Google Scholar] [CrossRef]
  353. Bould, J.; Clegg, W.; Kennedy, J.D.; Teat, S.J. Isomeric icosaboranes B20H26: The synchrotron structure of 1,1’-bis(nido-decaboranyl). Acta Cryst. C 2001, 57, 779–780. [Google Scholar] [CrossRef]
  354. Barrett, S.A.; Greenwood, N.N.; Kennedy, J.D.; Thornton-Pett, M. The chemistry of isomers of icosaborane(26): Crystal and molecular structure of 1,2’-bi(nido-decaboranyl). Polyhedron 1985, 4, 1981–1984. [Google Scholar] [CrossRef]
  355. Greenwood, N.N.; Kennedy, J.D.; McDonald, W.S.; Staves, J.; Taylorson, D. Isomers of B20H26: Structural characterisation by X-ray diffraction of 2,2’-bi(nido-decaboranyl). J. Chem. Soc. Chem. Commun. 1979, 17–18. [Google Scholar] [CrossRef]
  356. Boocock, S.K.; Greenwood, N.N.; Kennedy, J.D.; McDonald, W.S.; Staves, J. The chemistry of isomeric icosaboranes, B20H26. Molecular structures and physical characterization of 2,2’-bi(nido-decaboranyl) and 2,6’-bi(nido-decaboranyl). J. Chem. Soc. Dalton Trans. 1980, 790–796. [Google Scholar] [CrossRef]
  357. Boocock, S.K.; Greenwood, N.N.; Kennedy, J.D.; Taylorson, D. Isomers of B20H26: Elucidation of the structure of 6,6’-bi(nido-decaboranyl) by 11B-{1H} and 1H-{11B} n.m.r. spectroscopy. J. Chem. Soc. Chem. Commun. 1979, 16–17. [Google Scholar] [CrossRef]
  358. Bould, J.; Dörfler, U.; Clegg, W.; Teat, S.J.; Thornton-Pett, M.; Kennedy, J.D. Triple linking of the decaboranyl cluster. Structure of [(SMe2)2B10H10(B10H13)2] as determined by synchrotron X-ray diffraction analysis. Chem. Commun. 2001, 1788–1789. [Google Scholar] [CrossRef]
  359. Olsen, F.P.; Vasavada, R.C.; Hawthorne, M.F. The chemistry of n-B18H22 and i-B18H22. J. Am. Chem. Soc. 1968, 90, 3946–3951. [Google Scholar] [CrossRef]
  360. Li, Y.; Sneddon, L.G. Improved synthetic route to n-B18H22. Inorg. Chem. 2006, 45, 470–471. [Google Scholar] [CrossRef]
  361. Londesborough, M.G.S.; Hnyk, D.; Bould, J.; Bould, J.; Serrano-Andres, L.; Sauri, V.; Oliva, J.M.; Kubat, P.; Polivka, T.; Lang, K. Distinct photophysics of the isomers of B18H22 explained. Inorg. Chem. 2012, 51, 1471–1479. [Google Scholar] [CrossRef]
  362. Londesborough, M.G.S.; Dolanský, J.; Braborec, J.; Cerdán, L. Interaction of anti-B18H22 with light. In Handbook of Boron Science with Applications in Organometallics, Catalysis, Materials and Medicine; Hosmane, N.S., Eagling, R., Eds.; World Scientific Publishing: London, UK, 2019; Volume 3, pp. 115–136. [Google Scholar] [CrossRef]
  363. Cerdan, L.; Frances-Monerris, A.; Roca-Sanjuan, D.; Bould, J.; Dolandky, J.; Fuciman, M.; Londesborough, M.G.S. Unveiling the role of upper excited electronic states in the photochemistry and laser performance of anti-B18H22. J. Mater. Chem. C 2020, 8, 12806–12818. [Google Scholar] [CrossRef]
  364. Tan, C.; Zhang, B.; Chen, J.; Zhang, L.; Huang, X.; Meng, H. Study of hydrolysis kinetic of new laser material [anti-B18H22]. Russ. J. Inorg. Chem. 2019, 64, 1359–1364. [Google Scholar] [CrossRef]
  365. Ševčik, J.; Urbanek, P.; Hanulikova, B.; Čapkova, T.; Urbanek, M.; Antoš, J.; Londesborough, M.G.S.; Bould, J.; Ghasemi, B.; Petřkovsky, L.; et al. The photostability of novel boron hydride blue emitters in solution and polystyrene matrix. Materials 2021, 14, 589. [Google Scholar] [CrossRef] [PubMed]
  366. Capkova, T.; Hanulikova, B.; Sevcik, J.; Urbanek, P.; Antos, J.; Urbanek, M.; Kuritka, I. Incorporation of the new anti-octadecaborane laser dyes into thin polymer films: A temperature-dependent photoluminescence and infrared spectroscopy study. Int. J. Mol. Sci. 2022, 23, 8832. [Google Scholar] [CrossRef] [PubMed]
  367. Anderson, K.P.; Rheingold, A.L.; Djurovich, P.I.; Soman, O.; Spokoyny, A.M. Synthesis and luminescence of monohalogenated B18H22 clusters. Polyhedron 2022, 227, 116099. [Google Scholar] [CrossRef]
  368. Ehn, M.; Bavol, D.; Bould, J.; Strnad, V.; Litecká, M.; Lang, K.; Kirakci, K.; Clegg, W.; Waddell, P.G.; Londesborough, M.G.S. A window into the workings of anti-B18H22 luminescence—Blue-fluorescent isomeric pair 3,3’-Cl2-B18H20 and 3,4’-Cl2-B18H20 (and others). Molecules 2023, 28, 4505. [Google Scholar] [CrossRef]
  369. Anderson, K.P.; Waddington, M.A.; Balaich, G.J.; Stauber, J.M.; Bernier, N.A.; Caram, J.R.; Djurovich, P.I.; Spokoyny, A.M. A molecular boron cluster-based chromophore with dual emission. Dalton Trans. 2020, 49, 16245–16251. [Google Scholar] [CrossRef] [PubMed]
  370. Anderson, K.P.; Hua, A.S.; Plumley, J.B.; Ready, A.D.; Rheingold, A.L.; Peng, T.L.; Djurovich, I.; Kerestes, C.; Snyder, N.A.; Andrews, A.; et al. Benchmarking the dynamic luminescence properties and UV stability of B18H22-based materials. Dalton Trans. 2022, 51, 9223–9228. [Google Scholar] [CrossRef]
  371. Londesborough, M.G.S.; Dolansky, J.; Bould, J.; Braborec, J.; Kirakci, K.; Lang, K.; Cisařova, I.; Kubat, P.; Roca-Sanjuan, D.; Frances-Monerris, A.; et al. Effect of iodination on the photophysics of the laser borane anti-B18H22: Generation of efficient photosensitizers of oxygen. Inorg. Chem. 2019, 58, 10248–10259. [Google Scholar] [CrossRef]
  372. Anderson, K.P.; Djurovich, P.I.; Rubio, V.P.; Liang, A.; Spokoyny, A.M. Metal-catalyzed and metal-free nucleophilic substitution of 7-I-B18H21. Inorg. Chem. 2022, 61, 15051–15057. [Google Scholar] [CrossRef]
  373. Bould, J.; Lang, K.; Kirakci, K.; Cerdan, L.; Roca-Sanjuan, D.; Frances-Monerris, A.; Clegg, W.; Waddell, P.G.; Fuciman, M.; Polivka, T.; et al. A series of ultra-efficient blue borane fluorophores. Inorg. Chem. 2020, 59, 17058–17070. [Google Scholar] [CrossRef]
  374. Londesborough, M.G.S.; Lang, K.; Clegg, W.; Waddell, P.G.; Bould, J. Swollen polyhedral volume of the anti-B18H22 cluster via extensive methylation: Anti-B18H8Cl2Me12. Inorg. Chem. 2020, 59, 2651–2654. [Google Scholar] [CrossRef]
  375. Sauri, V.; Oliva, J.M.; Hnyk, D.; Bould, J.; Braborec, J.; Merchan, M.; Kubat, P.; Cisařova, I.; Lang, K.; Londesborough, M.G.S. Tuning the Photophysical Properties of anti-B18H22: Efficient intersystem crossing between excited ainglet and triplet states in new 4,4’-(HS)2-anti-B18H20. Inorg. Chem. 2013, 52, 9266–9274. [Google Scholar] [CrossRef]
  376. Londesborough, M.G.S.; Dolansky, J.; Cerdan, L.; Lang, K.; Jelinek, T.; Oliva, J.M.; Hnyk, D.; Roca-Sanjuan, D.; Frances-Monerris, A.; Martinčik, J.; et al. Thermochromic fluorescence from B18H20(NC5H5)2: An inorganic–organic composite luminescent compound with an unusual molecular geometry. Adv. Opt. Mater. 2017, 5, 1600694. [Google Scholar] [CrossRef]
  377. Londesborough, M.G.S.; Dolansky, J.; Jelinek, T.; Kennedy, J.D.; Cisařova, I.; Kennedy, R.D.; Roca-Sanjuan, D.; Frances-Monerris, A.; Lang, K.; Clegg, W. Substitution of the laser borane anti-B18H22 with pyridine: A structural and photophysical study of some unusually structured macropolyhedral boron hydrides. Dalton Trans. 2018, 47, 1709–1725. [Google Scholar] [CrossRef] [PubMed]
  378. Chen, J.; Xiong, L.; Zhang, L.; Huang, X.; Meng, H.; Tan, C. Synthesis, aggregation-induced emission of a new anti-B18H22-isoquinoline hybrid. Chem. Phys. Lett. 2020, 747, 137328. [Google Scholar] [CrossRef]
  379. Xiong, L.; Zheng, Y.; Wang, H.; Yan, J.; Huang, X.; Meng, H.; Tan, C. A novel AIEE active anti-B18H22 derivative-based Cu2+ and Fe3+ fluorescence off-on-off sensor. Methods Appl. Fluoresc. 2022, 10, 035004. [Google Scholar] [CrossRef]
  380. Jelinek, T.; Kennedy, J.D.; Štibr, B.; Thornton-Pett, M. Macropolyhedral boron-containing cluster chemistry. A reductive trimerisation of MeNC to give an imidazole-based carbene stabilized by coordination to boron in an eighteen-vertex cluster compound. J. Chem. Soc. Chem. Commun. 1994, 1999–2000. [Google Scholar] [CrossRef]
  381. Jelinek, T.; Kilner, C.A.; Štibr, B.; Thornton-Pett, M.; Kennedy, J.D. Macropolyhedral borane reaction chemistry: Reductive oligomerisation of terBuNC by anti-B18H22 to give the boron-coordinated {(terBuNHCH){terBuNHC(CN)}CH2:} carbene residue. Inorg. Chem. Commun. 2005, 8, 491–494. [Google Scholar] [CrossRef]
  382. Patel, D.K.; Sooraj, B.S.; Kirakci, K.; Macháček, J.; Kučeráková, M.; Bould, J.; Dušek, M.; Frey, M.; Neumann, C.; Ghosh, S.; et al. Macropolyhedral syn-B18H22, the “forgotten” isomer. J. Am. Chem. Soc. 2023, 145, 17975–17986. [Google Scholar] [CrossRef] [PubMed]
  383. Ehn, M.; Litecká, M.; Londesborough, M.G.S. Unexpected minor products from the thermal auto-fusion of arachno-SB8H12: Luminescent 4-(HS)-syn-B18H21 and 3-(HS)-syn-B18H21. Inorg. Chem. Commun. 2023, 155, 111021. [Google Scholar] [CrossRef]
  384. Jelínek, T.; Grüner, B.; Císařová, I.; Štíbr, B.; Kennedy, J.D. Macropolyhedral boron-containing cluster chemistry: The reaction of syn-B18H22 with SMe2 and I2 in monoglyme: Structure of [7-(SMe2)-syn-B18H20]. Inorg. Chem. Commun. 2007, 10, 125–128. [Google Scholar] [CrossRef]
Figure 1. Structure and numbering of atoms in decaborane B10H14.
Figure 1. Structure and numbering of atoms in decaborane B10H14.
Molecules 28 06287 g001
Scheme 1. Synthesis of decaborane(14) from sodium tetrahydroborate NaBH4.
Scheme 1. Synthesis of decaborane(14) from sodium tetrahydroborate NaBH4.
Molecules 28 06287 sch001
Scheme 2. Preparation of the [B10H13] anion and its tautomeric forms.
Scheme 2. Preparation of the [B10H13] anion and its tautomeric forms.
Molecules 28 06287 sch002
Figure 2. Solid state structures of the (HPS)+ cation (left) and of the [nido-B10H13] anion (right) in the crystal structure of (HPS)[B10H13].
Figure 2. Solid state structures of the (HPS)+ cation (left) and of the [nido-B10H13] anion (right) in the crystal structure of (HPS)[B10H13].
Molecules 28 06287 g002
Scheme 3. Preparation of the [arachno-B10H14]2− anion.
Scheme 3. Preparation of the [arachno-B10H14]2− anion.
Molecules 28 06287 sch003
Figure 3. Solid state structure of the 2-iodo derivative of decaborane [2-I-B10H13].
Figure 3. Solid state structure of the 2-iodo derivative of decaborane [2-I-B10H13].
Molecules 28 06287 g003
Figure 4. Solid state structures of [5-I-B10H13] (left) and [6-I-B10H13] (right).
Figure 4. Solid state structures of [5-I-B10H13] (left) and [6-I-B10H13] (right).
Molecules 28 06287 g004
Figure 5. Solid state structures of [5-Br-B10H13] (left) and [6-Br-B10H13] (right).
Figure 5. Solid state structures of [5-Br-B10H13] (left) and [6-Br-B10H13] (right).
Molecules 28 06287 g005
Figure 6. Solid state structure of the 5,10-dibromo derivative of decaborane [5,10-Br2-B10H12].
Figure 6. Solid state structure of the 5,10-dibromo derivative of decaborane [5,10-Br2-B10H12].
Molecules 28 06287 g006
Figure 7. Solid state structures of [6-Cl-B10H13] (top) as well as the [6-Cl-B10H12] (bottom left) and [5-Cl-B10H12] (bottom right) anions in the crystal structures of the corresponding protonated Proton Sponge salts.
Figure 7. Solid state structures of [6-Cl-B10H13] (top) as well as the [6-Cl-B10H12] (bottom left) and [5-Cl-B10H12] (bottom right) anions in the crystal structures of the corresponding protonated Proton Sponge salts.
Molecules 28 06287 g007
Figure 8. Solid state structure of the 6-fluoro derivative of decaborane [6-F-B10H13].
Figure 8. Solid state structure of the 6-fluoro derivative of decaborane [6-F-B10H13].
Molecules 28 06287 g008
Figure 9. Solid state structures of [5-MeO-B10H13] (top left), [6-t-BuO-B10H13] (top right), and [μ-6,6′-(OC6H10O)-(B10H13)2] (bottom).
Figure 9. Solid state structures of [5-MeO-B10H13] (top left), [6-t-BuO-B10H13] (top right), and [μ-6,6′-(OC6H10O)-(B10H13)2] (bottom).
Molecules 28 06287 g009
Figure 10. Solid state structures of [6-TfO-B10H13] (left) and [5-TfO-B10H13] (right).
Figure 10. Solid state structures of [6-TfO-B10H13] (left) and [5-TfO-B10H13] (right).
Molecules 28 06287 g010
Figure 11. Structure of the [6-arachno-AcO-B10H13]2− anion in the crystal structure of (C2mim)2[6-AcO-B10H13].
Figure 11. Structure of the [6-arachno-AcO-B10H13]2− anion in the crystal structure of (C2mim)2[6-AcO-B10H13].
Molecules 28 06287 g011
Figure 12. Solid state structures of the mercapto derivatives of decaborane [1-HS-B10H13] (left), [2-HS-B10H13] (middle), and [1,2-(HS)2-B10H12] (right).
Figure 12. Solid state structures of the mercapto derivatives of decaborane [1-HS-B10H13] (left), [2-HS-B10H13] (middle), and [1,2-(HS)2-B10H12] (right).
Molecules 28 06287 g012
Figure 13. Solid state structure of [nido-5-Me2S-B10H12].
Figure 13. Solid state structure of [nido-5-Me2S-B10H12].
Molecules 28 06287 g013
Figure 14. Solid state structure of [arachno-6,9-(MeC≡N)2-B10H12].
Figure 14. Solid state structure of [arachno-6,9-(MeC≡N)2-B10H12].
Molecules 28 06287 g014
Figure 15. Solid state structures of the decaborane-based amidines [6,9-(Bu2N(Me)C=HN)2-B10H12] (top) and [6,9-(PhHN(Me)C=HN)2-B10H12] (bottom).
Figure 15. Solid state structures of the decaborane-based amidines [6,9-(Bu2N(Me)C=HN)2-B10H12] (top) and [6,9-(PhHN(Me)C=HN)2-B10H12] (bottom).
Molecules 28 06287 g015
Figure 16. Solid state structure of [arachno-6,9-(H3N)2-B10H12].
Figure 16. Solid state structure of [arachno-6,9-(H3N)2-B10H12].
Molecules 28 06287 g016
Scheme 4. Synthesis of 6,9-bis(pyridinium) derivatives [6,9-L2-B10H12].
Scheme 4. Synthesis of 6,9-bis(pyridinium) derivatives [6,9-L2-B10H12].
Molecules 28 06287 sch004
Figure 17. Solid state structures of [arachno-6,9-(HC≡C-o-C5H4N)2-B10H12] (left) and [arachno-6,9-(N≡C-o-C5H4N)2-B10H12] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 17. Solid state structures of [arachno-6,9-(HC≡C-o-C5H4N)2-B10H12] (left) and [arachno-6,9-(N≡C-o-C5H4N)2-B10H12] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g017
Scheme 5. Synthesis of [arachno-6,6-Py2-B10H12] and its transformation to [arachno-6,9-Py2-B10H12].
Scheme 5. Synthesis of [arachno-6,6-Py2-B10H12] and its transformation to [arachno-6,9-Py2-B10H12].
Molecules 28 06287 sch005
Figure 18. Solid state structures of [μ-6,6′-pyrazine-(9-Me2S-B10H12)2] (top) and [6,9-(NC5H4C5H4N)2-B10H12] (bottom).
Figure 18. Solid state structures of [μ-6,6′-pyrazine-(9-Me2S-B10H12)2] (top) and [6,9-(NC5H4C5H4N)2-B10H12] (bottom).
Molecules 28 06287 g018
Scheme 6. Reactions of [arachno-6,9-(Me2S)2-B10H12] with 1,4-bis[β-(4-pyridyl)vinyl]benzene and 1,4-bis[β-(4-quinolyl)vinyl]benzene.
Scheme 6. Reactions of [arachno-6,9-(Me2S)2-B10H12] with 1,4-bis[β-(4-pyridyl)vinyl]benzene and 1,4-bis[β-(4-quinolyl)vinyl]benzene.
Molecules 28 06287 sch006
Figure 19. Solid state structures of [6,9-(HIm)2-Me2S-B10H12] (top left), [6,9-(MeIm)2-Me2S-B10H12] (top right), [6,9-(EtIm)2-Me2S-B10H12] (bottom left), and [6,9-(BuIm)2-Me2S-B10H12] (top right). Hydrogen atoms of alkyl groups are omitted for clarity.
Figure 19. Solid state structures of [6,9-(HIm)2-Me2S-B10H12] (top left), [6,9-(MeIm)2-Me2S-B10H12] (top right), [6,9-(EtIm)2-Me2S-B10H12] (bottom left), and [6,9-(BuIm)2-Me2S-B10H12] (top right). Hydrogen atoms of alkyl groups are omitted for clarity.
Molecules 28 06287 g019
Figure 20. Solid state structures of [6-SCN-B10H13] (left) and [6-N3-μ-5,6-NH2-B10H11] (right).
Figure 20. Solid state structures of [6-SCN-B10H13] (left) and [6-N3-μ-5,6-NH2-B10H11] (right).
Molecules 28 06287 g020
Figure 21. Solid state structures of exo,endo-[6,9-(PhMe2P)2-arachno-B10H12] (left) and exo,exo-[6,9-(PhMe2P)2-arachno-B10H12] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 21. Solid state structures of exo,endo-[6,9-(PhMe2P)2-arachno-B10H12] (left) and exo,exo-[6,9-(PhMe2P)2-arachno-B10H12] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g021
Figure 22. Solid state structure of exo,endo-[6,9-(PhMe2P)2-2-Br-arachno-B10H12]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 22. Solid state structure of exo,endo-[6,9-(PhMe2P)2-2-Br-arachno-B10H12]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g022
Figure 23. Solid state structure of the [arachno-μ-6,9-Ph2P-B10H12] anion. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 23. Solid state structure of the [arachno-μ-6,9-Ph2P-B10H12] anion. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g023
Figure 24. Solid state structures of [1-I-2,3,4,5,6,7,8-Me7-B10H6] (left) and [1-TfO-2,3,4,5,6,7,8-Me7-B10H6] (right). Hydrogen atoms of methyl groups are omitted for clarity.
Figure 24. Solid state structures of [1-I-2,3,4,5,6,7,8-Me7-B10H6] (left) and [1-TfO-2,3,4,5,6,7,8-Me7-B10H6] (right). Hydrogen atoms of methyl groups are omitted for clarity.
Molecules 28 06287 g024
Figure 25. Solid state structures of [6-Chx-8-Me2S-B10H11] (left) and [6-Thx-8-Me2S-B10H11] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 25. Solid state structures of [6-Chx-8-Me2S-B10H11] (left) and [6-Thx-8-Me2S-B10H11] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g025
Figure 26. Solid state structures of [6-Thx-B10H13] (top left), [6-MeC(O)CH2CH2CH2CH2-B10H13] (top right), [6-Me3SiCH2CH2CH2-B10H13] (bottom left), and [6-H2C=CHCH2SiMe2CH2CH2CH2-B10H13] (bottom right). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 26. Solid state structures of [6-Thx-B10H13] (top left), [6-MeC(O)CH2CH2CH2CH2-B10H13] (top right), [6-Me3SiCH2CH2CH2-B10H13] (bottom left), and [6-H2C=CHCH2SiMe2CH2CH2CH2-B10H13] (bottom right). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g026
Figure 27. Solid state structures of [nido-6-C6H11-B10H13] (top), [6-(4′-cyclohexenyl)-B10H13] (bottom left) and [6-(5′-norbornenyl)-B10H13] (bottom right).
Figure 27. Solid state structures of [nido-6-C6H11-B10H13] (top), [6-(4′-cyclohexenyl)-B10H13] (bottom left) and [6-(5′-norbornenyl)-B10H13] (bottom right).
Molecules 28 06287 g027
Figure 28. Solid state structures of [μ-6,6′-(CH2)6-(B10H13)2] (top), [μ-6,6′-(2″,5″-norbornyl)-(B10H13)2] (middle), and [μ4-6,6′,6″,6‴-Si-(6-(CH2)3-B10H13)4] (bottom). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 28. Solid state structures of [μ-6,6′-(CH2)6-(B10H13)2] (top), [μ-6,6′-(2″,5″-norbornyl)-(B10H13)2] (middle), and [μ4-6,6′,6″,6‴-Si-(6-(CH2)3-B10H13)4] (bottom). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g028
Figure 29. Solid state structure of [6,9-(C5H11)2-5-I-B10H11]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 29. Solid state structure of [6,9-(C5H11)2-5-I-B10H11]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g029
Figure 30. Solid state structures of [6,9-((E)-Br(CH2)2CH=CH)2-B10H12] (left) and [6,9-((E)-Me3SiCH=CH)2-B10H12] (right).
Figure 30. Solid state structures of [6,9-((E)-Br(CH2)2CH=CH)2-B10H12] (left) and [6,9-((E)-Me3SiCH=CH)2-B10H12] (right).
Molecules 28 06287 g030
Figure 31. Solid state structure of [6,9-(c-C6H11CH2(H2C=)C)2-B10H12]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 31. Solid state structure of [6,9-(c-C6H11CH2(H2C=)C)2-B10H12]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g031
Figure 32. Solid state structure of [6-Me3Si(CH2)3-9-(E)-m-HC≡CC6H4CH=CH-B10H12].
Figure 32. Solid state structure of [6-Me3Si(CH2)3-9-(E)-m-HC≡CC6H4CH=CH-B10H12].
Molecules 28 06287 g032
Figure 33. Solid state structures of [9,9′-μ-CH2CH2-(6-C5H11-B10H12)2] (top) and [9,9′-μ-CH2CH2-(6-Me3Si(CH2)3-B10H12)2] (bottom).
Figure 33. Solid state structures of [9,9′-μ-CH2CH2-(6-C5H11-B10H12)2] (top) and [9,9′-μ-CH2CH2-(6-Me3Si(CH2)3-B10H12)2] (bottom).
Molecules 28 06287 g033
Figure 34. Solid state structure of [9,9′-μ-CH=CH-(6-Me3Si(CH2)3-B10H12)2].
Figure 34. Solid state structure of [9,9′-μ-CH=CH-(6-Me3Si(CH2)3-B10H12)2].
Molecules 28 06287 g034
Figure 35. Solid state structures of [6-Me3Si(Me)C=CH-5-Me2S-B10H11] (top left), [6-Me3Si(Bu)C=CH-5-Me2S-B10H11] (top right), and [6-(Me3Si)2C=CH-5-Me2S-B10H11] (bottom). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 35. Solid state structures of [6-Me3Si(Me)C=CH-5-Me2S-B10H11] (top left), [6-Me3Si(Bu)C=CH-5-Me2S-B10H11] (top right), and [6-(Me3Si)2C=CH-5-Me2S-B10H11] (bottom). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g035
Figure 36. Solid state structure of [μ-6(C),6′(C),5′(P)-C(t-Bu)PH-(nido-8-Me2S-B10H11)(nido-B10H12)]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 36. Solid state structure of [μ-6(C),6′(C),5′(P)-C(t-Bu)PH-(nido-8-Me2S-B10H11)(nido-B10H12)]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g036
Figure 37. Solid state structures of [μ-(exo-6(C),endo-6(N)-CH=CH-o-C5H4N)-9(N)-HC≡C-o-C5H4N-arachno-B10H11] (left) and [μ-(exo-6(C),endo-6(N)-(closo-1′,2′-C2B10H10-2′-)-o-C5H4N)-μ-(exo-8(C),exo-9(N)-CH2CH2-o-C5H4N)-arachno-B10H10] (right). Hydrogen atoms of organic substituents in the left structure are omitted for clarity.
Figure 37. Solid state structures of [μ-(exo-6(C),endo-6(N)-CH=CH-o-C5H4N)-9(N)-HC≡C-o-C5H4N-arachno-B10H11] (left) and [μ-(exo-6(C),endo-6(N)-(closo-1′,2′-C2B10H10-2′-)-o-C5H4N)-μ-(exo-8(C),exo-9(N)-CH2CH2-o-C5H4N)-arachno-B10H10] (right). Hydrogen atoms of organic substituents in the left structure are omitted for clarity.
Molecules 28 06287 g037
Figure 38. Solid state structures of [6-Ph-B10H13] (top left), [6-p-Tol-B10H13] (top right), and [6-(2′,4′,6′-iPr3-C6H2-B10H13] (bottom). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 38. Solid state structures of [6-Ph-B10H13] (top left), [6-p-Tol-B10H13] (top right), and [6-(2′,4′,6′-iPr3-C6H2-B10H13] (bottom). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g038
Figure 39. Solid state structures of [5-Ph-B10H13] (left) and [5,8-Ph2-B10H12] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Figure 39. Solid state structures of [5-Ph-B10H13] (left) and [5,8-Ph2-B10H12] (right). Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g039
Figure 40. Solid state structures of the [arachno-μ-6(C),9(N)-MeC=NH-B10H12] (left) and [arachno-endo-6-N≡C-B10H12]2− (right) anions.
Figure 40. Solid state structures of the [arachno-μ-6(C),9(N)-MeC=NH-B10H12] (left) and [arachno-endo-6-N≡C-B10H12]2− (right) anions.
Molecules 28 06287 g040
Figure 41. Solid state structures of [nido-6-(t-BuNH)(C5H6Me4NH)B-B10H13] (left) and [μ-5,6-(9-BBN)-B10H13] (right). Hydrogen atoms of alkyl groups are omitted for clarity.
Figure 41. Solid state structures of [nido-6-(t-BuNH)(C5H6Me4NH)B-B10H13] (left) and [μ-5,6-(9-BBN)-B10H13] (right). Hydrogen atoms of alkyl groups are omitted for clarity.
Molecules 28 06287 g041
Figure 42. Solid state structures of [arachno-1-(6′-nido-B10H13)-6,9-(Me2S)2-B10H11] (left) and [nido-4-(2′-nido-B10H13)-5-Me2S-B10H11] (right). Hydrogen atoms of alkyl groups are omitted for clarity.
Figure 42. Solid state structures of [arachno-1-(6′-nido-B10H13)-6,9-(Me2S)2-B10H11] (left) and [nido-4-(2′-nido-B10H13)-5-Me2S-B10H11] (right). Hydrogen atoms of alkyl groups are omitted for clarity.
Molecules 28 06287 g042
Figure 43. Solid state structure of [arachno-1,5-(6′-nido-B10H13)2-6,9-(Me2S)2-B10H10]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 43. Solid state structure of [arachno-1,5-(6′-nido-B10H13)2-6,9-(Me2S)2-B10H10]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g043
Figure 44. Structure of 5-(nido-pentaboran-2-yl)-6-(nido-pentaboran-1-yl)-nido-decaborane.
Figure 44. Structure of 5-(nido-pentaboran-2-yl)-6-(nido-pentaboran-1-yl)-nido-decaborane.
Molecules 28 06287 g044
Figure 45. Structures and numbering of atoms in syn- (left) and anti- (right) isomers of [B18H22]. Reprinted with permission from Ref. [359]. Copyright (1968) the American Chemical Society.
Figure 45. Structures and numbering of atoms in syn- (left) and anti- (right) isomers of [B18H22]. Reprinted with permission from Ref. [359]. Copyright (1968) the American Chemical Society.
Molecules 28 06287 g045
Figure 46. Solid state structures of anti-[7-Cl-B18H21] (top left), anti-[3,1′-Cl2-B18H20] (top right), anti-[4,4′-Cl2-B18H20] (middle left), anti-[3,3′-Cl2-B18H20] (middle right), and anti-[3,4′-Cl2-B18H20] (bottom).
Figure 46. Solid state structures of anti-[7-Cl-B18H21] (top left), anti-[3,1′-Cl2-B18H20] (top right), anti-[4,4′-Cl2-B18H20] (middle left), anti-[3,3′-Cl2-B18H20] (middle right), and anti-[3,4′-Cl2-B18H20] (bottom).
Molecules 28 06287 g046aMolecules 28 06287 g046b
Figure 47. Solid state structures of anti-[4-Br-B18H21] (left) and anti-[4,4′-Br2-B18H20] (right).
Figure 47. Solid state structures of anti-[4-Br-B18H21] (left) and anti-[4,4′-Br2-B18H20] (right).
Molecules 28 06287 g047
Figure 48. Solid state structures of anti-[7-I-B18H21] (left) and anti-[4,4′-I2-B18H20] (right).
Figure 48. Solid state structures of anti-[7-I-B18H21] (left) and anti-[4,4′-I2-B18H20] (right).
Molecules 28 06287 g048
Figure 49. Solid state structures of anti-[4,4′-Me2-B18H20] (top left), anti-[3,4,4′-Me3-B18H19] (top right), anti-[3,3′,4,4′-Me4-B18H18] (bottom left), and anti-[3,3′,4,4′-Et4-B18H18] (bottom right).
Figure 49. Solid state structures of anti-[4,4′-Me2-B18H20] (top left), anti-[3,4,4′-Me3-B18H19] (top right), anti-[3,3′,4,4′-Me4-B18H18] (bottom left), and anti-[3,3′,4,4′-Et4-B18H18] (bottom right).
Molecules 28 06287 g049aMolecules 28 06287 g049b
Figure 50. Solid state structure of anti-[2,2′-Cl2-1,1′,3,3′,4,4′,7,7′,8,8′,10,10′-Me12-B18H8]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 50. Solid state structure of anti-[2,2′-Cl2-1,1′,3,3′,4,4′,7,7′,8,8′,10,10′-Me12-B18H8]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g050
Figure 51. Solid state structure of anti-[4,4′-(HS)2-B18H20].
Figure 51. Solid state structure of anti-[4,4′-(HS)2-B18H20].
Molecules 28 06287 g051
Figure 52. Solid state structures of nido-arachno-[6′,9′-Py2-B18H20] (left) and anti-[8-Py-B18H21] (right).
Figure 52. Solid state structures of nido-arachno-[6′,9′-Py2-B18H20] (left) and anti-[8-Py-B18H21] (right).
Molecules 28 06287 g052
Figure 53. Solid state structure of anti-[7-{(MeNH)C3N2HMe2}-B18H20]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 53. Solid state structure of anti-[7-{(MeNH)C3N2HMe2}-B18H20]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g053
Figure 54. Solid state structure of anti-[7-{(t-BuNHCH){t-BuNHC(CN)}CH2}-B18H20]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 54. Solid state structure of anti-[7-{(t-BuNHCH){t-BuNHC(CN)}CH2}-B18H20]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g054
Figure 55. Solid state structures of syn-[1-HS-B18H21] (top left), syn-[3-HS-B18H21] (top right), and syn-[4-HS-B18H21] (bottom).
Figure 55. Solid state structures of syn-[1-HS-B18H21] (top left), syn-[3-HS-B18H21] (top right), and syn-[4-HS-B18H21] (bottom).
Molecules 28 06287 g055
Figure 56. Solid state structure of syn-[7-Me2S-B18H20]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 56. Solid state structure of syn-[7-Me2S-B18H20]. Hydrogen atoms of organic substituents are omitted for clarity.
Molecules 28 06287 g056
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sivaev, I.B. Decaborane: From Alfred Stock and Rocket Fuel Projects to Nowadays. Molecules 2023, 28, 6287. https://doi.org/10.3390/molecules28176287

AMA Style

Sivaev IB. Decaborane: From Alfred Stock and Rocket Fuel Projects to Nowadays. Molecules. 2023; 28(17):6287. https://doi.org/10.3390/molecules28176287

Chicago/Turabian Style

Sivaev, Igor B. 2023. "Decaborane: From Alfred Stock and Rocket Fuel Projects to Nowadays" Molecules 28, no. 17: 6287. https://doi.org/10.3390/molecules28176287

Article Metrics

Back to TopTop