Next Article in Journal
Influence of Dispersed Phase Content on the Mechanical Properties of Electroless Nanocomposite Ni-P/Si3N4 and Hybrid Ni-P/Si3N4/Graphite Layers Deposited on the AW-7075 Alloy
Next Article in Special Issue
Transition Metal Borides for All-in-One Radiation Shielding
Previous Article in Journal
Critical Evaluation of the Methods for the Characterization of the Degree of Sulfonation for Electron Beam Irradiated and Non-Irradiated Sulfonated Poly(ether ether ketone) Membranes
Previous Article in Special Issue
Exploring the Structural, Electronic, Magnetic, and Transport Properties of 2D Cr, Fe, and Zr Monoborides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Composites and Materials Prepared from Boron Cluster Anions and Carboranes

by
Varvara V. Avdeeva
1,*,
Svetlana E. Nikiforova
1,
Elena A. Malinina
1,
Igor B. Sivaev
2,3 and
Nikolay T. Kuznetsov
1
1
Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, 31 Leninskii Av., Moscow 119991, Russia
2
Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilov Str., Moscow 119991, Russia
3
Basic Department of Chemistry of Innovative Materials and Technologies, Plekhanov Russian University of Economics, 36 Stremyannyi Line, Moscow 117997, Russia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(18), 6099; https://doi.org/10.3390/ma16186099
Submission received: 4 August 2023 / Revised: 4 September 2023 / Accepted: 4 September 2023 / Published: 6 September 2023
(This article belongs to the Special Issue Development of Boron-Based Materials)

Abstract

:
Here, we present composites and materials that can be prepared starting with boron hydride cluster compounds (decaborane, decahydro-closo-decaborate and dodecahydro-closo-dodecaborate anions and carboranes). Recent examples of their utilization as boron protective coatings including using them to synthesize boron carbide, boron nitride, metal borides, metal-containing composites, and neutron shielding materials are discussed. The data are generalized demonstrate the versatile application of materials based on boron cluster anions and carboranes in various fields.

Graphical Abstract

1. Introduction

The driving force behind the intensive development of the chemistry of boron hydride cluster compounds in the 1950s was their use as high-energy materials [1,2,3]. This topic is still discussed, although it attracts less attention [4,5,6,7,8,9,10,11,12,13,14]. Currently, the most attention is paid to the use of boron cluster compounds in medicine [15,16,17,18,19,20,21,22,23,24,25,26,27]. As for the use of boron clusters in materials science, complex molecular and supramolecular structures, such as nanocars and nanotrains [28,29,30,31], molecular machines, and switches [32,33,34,35,36], MOFs [37,38,39,40,41,42,43,44,45,46], etc., attract the most attention. However, there is another less spectacular but no less important direction of research on the use of boron cluster compounds in the science of materials. This direction of research consists of the thermal decomposition of various boron clusters with the formation of boron-containing coatings and ceramic materials, which is somewhat reminiscent of their use as high-energy materials, since in this case a complex molecular structure is also converted into simple molecules.
In this contribution to the field, we present an attempt to consider the use of boron clusters to obtain various boron-containing materials, including boron-containing coatings and films, nanostructured boron carbide and metal borides, and others. The most readily available decaborane(14) B10H14, ortho- and meta-carboranes 1,2-C2B10H12 and 1,2-C2B10H12, decahydro-closo-decaborate [B10H10]2− and octadecahydro-conjucto-eicosaborate [trans-B20H18]2− anions (Figure 1) are considered as the boron clusters. Decaborane(14) can be readily prepared in a two-step procedure from sodium tetrahydroborate NaBH4 [47]. The well-known ortho-carborane is obtained by introducing the acetylene molecule into the open boron cage of decaborane(14), and its thermal isomerization leads to meta-carborane [48]. The decahydro-closo-decaborate anion [49,50,51] is formed by heating decaborane(14) in the presence of triethylamine, and its mild oxidation results in the octadecahydro-conjucto-eicosaborate anion [52,53].

2. Decaborane as Boron Source for Boron-Containing Materials

Metal Organic Chemical Vapor Deposition (MOCVD) is widely used for creating high-purity crystalline semiconducting thin films and micro/nano structures for microelectronics [54,55,56,57]. Therefore, it is not surprising that volatile boron hydrides such as decaborane have been proposed for the preparation of various boron-containing coatings. At first glance, it may seem that for these purposes it is more convenient to use other more available and cheap volatile boron compounds, such as BCl3, BF3 or diborane. However, one drawback of MOCVD is the aggressive, toxic and explosive nature of the precursor gases, which makes them difficult to use in small research laboratories. This fully applies to both aggressive and corrosive boron halides and highly toxic and flammable diborane. Therefore, decaborane B10H14, despite its higher cost and toxicity [58,59], is in many respects a more convenient source of boron for these purposes. Furthermore, decaborane has advantages as a source material for boron coating because high-purity decaborane is easy to obtain with a sublimation purification process.
Amorphous boron films of 0.1–1.5 μm thickness have been prepared on sapphire, silicon, and tantalum as substrates by the pyrolysis of decaborane in the molecular flow region (≤10−4 torr) and in a temperature range of 350–1200 °C. It is found that the deposition rate of the boron films is proportional to the decaborane partial pressure and the substrate temperature. The electrical conductivities vary from 3 × 10−5 S cm−1 at 77 K to 30 S cm−1 at 1000 K, and the activation energy is 1.07 eV in the intrinsic temperature range (700–1000 K). The maximum value of thermoelectric power is about 420 μV deg−1 at 700 K, and its polarity is positive between 500 and 1000 K [60]. The preparation of polycrystalline α-rhombohedral boron films by pyrolysis of decaborane has also been reported [61].
Boron coating also can be obtained by the plasma-assisted enhanced chemical vapor deposition (PECVD) of decaborane [62,63]. In particular, decaborane has been proposed as a boron source for the boronization of JT-60U Tokamak to reduce the influx of impurities during plasma discharge [64]. The deposition of boron films on polished p-type Si(111) surface by synchrotron-radiation-induced chemical vapor deposition (SR-CVD) of decaborane was reported [65].
When using ammonia or dinitrogen as additives, the chemical vapor deposition of decaborane can be used to obtain thin films of boron nitride [66], nanosheets [67], and nanotubes [68], competing with borazine. In particular, boron nitride nanotubes (BNNTs) grown at 1200–1300 °C from decaborane were double- and multiwalled, with the double-walled nanotubes having ~2 nm inner diameters and the multiwalled nanotubes (~10 walls) having ~4–5 nm inner diameters and ~12–14 nm outer diameters. The nanotubes grown at 1300 °C were longer, averaging ~0.6 μm, whereas those grown at 1200 °C had average lengths of ~0.2 μm [68].
The pyrolysis of decaborane can also be used to prepare boron nanoparticles [69] and microcrystals [70,71]. α-Tetragonal boron crystals were obtained at a pressure of 8–9 GPa and temperatures in the range 1100–1600 °C, while β-rhombohedral boron crystals grow at 3 GPa and 1200 °C [70]. The α-tetragonal boron crystals synthesized demonstrate semiconducting properties of conductivity with the energy gap Eg ≈ 1.5 eV [71].
The chemical vapor deposition of decaborane was also used to prepare various metal boride thin films, including nickel [72], strontium [73], gadolinium [74], neodymium [75], and ytterbium [76,77].
A synthetic route to metal borides TiB2, ZrBz2, HfB2, NbB2, and TaB2 by heating the decaborane-pimelonitrilium polymer [-6-B10H12-(NC(CH2)5CN)]n- and the corresponding finely dispersed metal oxides above 1400 °C was proposed. The metal boride powders were found to be highly crystalline, with grain sizes dependent on processing temperatures [78].
Various boron-carbide-containing materials were prepared using various decaborane-based single-molecular precursors, such as [μ-6,6′-(CH2)6-(B10H13)2] [79,80], [μ-6,6′-(1′,5′-cyclooctyl)-(B10H13)2] [81], [μ-6,6′-(2′,5′-norbornenyl)-(B10H13)2] [81], or polymers, including [-6-B10H12-Ph2POPPh2]n- [82], [-6-B10H12-(CH2)6]n- [83,84], [-6-B10H12-(2′,5′-norbornenyl)]n- [81,83,84,85,86,87,88,89], [-6-B10H12-(1′,5′-cyclooctenyl)]n- [81,83], and [-6-B10H12-(1′,4′-cyclooctenyl)]n- [81,83].
In particular, [μ-6,6′-(CH2)6-(B10H13)2] (Figure 2) appears to be an ideal precursor for the synthesis of boron carbide nanofibers (Figure 3) using the templating technique: (i) it is readily synthesized in large amounts using the Ti-catalyzed reaction [90]; (ii) it contains no other ceramic-forming elements and has a desirable boron-to-carbon ratio, thus yielding boron-rich boron carbide compositions upon pyrolysis; (iii) it is stable as a liquid, allowing it to be absorbed into the membrane without decomposition; and (iv) upon pyrolysis, it undergoes a cross-linking reaction at relatively low temperatures (220 °C), which slows the loss of material by volatilization, thereby generating high ceramic and chemical yields [79].
The bis(decaboranyl)-hexane precursor [μ-6,6′-(CH2)6-(B10H13)2] can also be used for the preparation of ordered mesoporous boron carbide materials with high specific surface areas up to 778 m2/g and hexagonal pore arrangement symmetries [80].
It should be noted that an alternative possibility of using pentaborane B5H9 instead of decaborane to obtain boron carbide compositions was previously considered [91,92,93]; however, after the destruction of pentaborane stocks stored since the 1960s [94], this aim was abandoned.
The B4C/BN-containing ceramic materials can be prepared using the pyrolysis of polymeric Lewis base adducts of decaborane [-6-B10H12-(diamine)]n- (diamine is ethylenediamine, 1,1-dimethylethylenediamine, 1,1,2,2-tetramethylethylenediamine) [95,96]. In particular, the [-6-B10H12-(ethylenediamine)]n-polymer fibers upon pyrolysis at 1000 °C in an argon atmosphere retain their shape and give black ceramic fibers with a diameter of 3 to 5 μm, which have a round shape, a smooth surface, and no obvious major flaws (Figure 4). Other [-6-B10H12-(diamine)]n- polymers were capable of forming fibers. The polymers derived from 1,1,2,2-tetramethylethylenediamine and from the 85/15 1,1-dimethylethylenediamine/1,1,2,2-tetramethylethylenediamine mixture melt when heated (mp 246–250 °C and 222–225 °C, respectively) and may be suitable for melt-spinning [87].
Low-crystalline boron nitride was prepared by the reaction of triammoniadecaborane B10H14·3NH3 and hydrazine or ammonia at 125 MPa and 650–700 °C. The prepared low-crystalline boron nitride passed into cubic boron nitride at 1200–1300 °C and 6.5 GPa in the presence of 20 mol.% AINas a catalyst [97,98].

3. Carborane as a Boron Source for Boron-Containing Materials

Chemical vapor deposition methods are widely used for the manufacture of boron carbide films due to the better controlled deposition process and the high-quality boron carbide production. Mixtures of boron trichloride, methane, and hydrogen are usually used for the CVD of boron carbide films. Since chlorides are highly dangerous and the synthesis process requires high temperatures, the replacement of BCl3 with organoboranes has become a trend in recent years. At first it seemed that small organoboron molecules such as trimethylboron and triethylboron could be a good alternative, but they proved to be overly reactive. Taking into account that ortho-carborane [H2C2B10H10] provides a suitable ratio of B and C from a single molecular source, it seems to be an attractive source for preparing boron carbide materials.
Semiconducting boron carbide represents a new class of materials with potential applications in neutron detection because 10B has a high cross-section (approximately 3800 barns) for neutrons at lower energies (~25 meV), based on the 10B(n,α)7Li neutron capture reaction [99,100,101,102,103,104,105]. This aroused great interest in the fabrication of boron carbide films using the PECVD of ortho-carborane, and the effect of the process parameters, such as temperature and total pressure, on the composition, microstructure, morphology, and properties of the boron carbide films obtained were studied [106,107,108,109,110,111,112,113,114]. In particular, the boron carbide film prepared at low temperatures and pressures (Tdep = 900 °C and Ptot = 100 Pa) showed a comparatively flat morphology, whereas the boron carbide films prepared at low temperature and high pressure (Tdep = 900 °C and Ptot = 50,000 Pa) appeared as round bulges. The boron carbide films prepared at a high temperature and relatively low pressure (Tdep = 1100 °C and Ptot = 5000 Pa) exhibited a cauliflower-like surface, while the films prepared at high temperature and high pressure (Tdep = 1200 °C and Ptot = 50,000 Pa) exhibited a uniform granular surface (Figure 5) [112].
Semiconducting boron carbide films can be also prepared through the PECVD of meta-carborane, which differs from ortho-carborane only in the arrangement of carbon atoms in the icosahedral cage [115,116,117]. It was found that meta-carborane and ortho-carborane form self-doped n-type and p-type boron-carbides, respectively [115,116].
It was shown that neutron detectors and neutron voltaic devices, based on semiconducting boron carbides, contrary to most other electrical devices, may improve with some radiation exposure and are robust against radiation-induced device degradation and failure [104]. The main causes for the poor neutron detection device performance are the insufficiently thick depletion region of the device, the need for a thicker device to come closer to neutron opacity, and the need for better charge collection while maintaining low reverse bias leakage currents [118]. It was found that the semiconducting boron carbide prepared by PECVD of composites of ortho- and meta-carboranes and aromatic or heteroaromatic compounds [119,120,121,122,123,124,125] demonstrate improvements in both charge collection and reverse bias leakage currents, which is attributed to an increase in the hole carrier lifetimes.
The introduction of metallocenes Cp2M (M = Ni, Co, Fe, Mn) together with ortho- or meta-carboranes during the PECVD process results in the corresponding transition metal doping of semiconducting boron carbide films [126,127,128,129,130,131].
Another important area of using ortho- and meta-carboranes to create protective boron carbide coatings is the boronization of tokamaks. The plasma-chemical deposition of a protective coating on the first wall of a fusion device using a chemically active gas (precursor) remains to date one of the primary ways to protect plasma against cooling impurities. This method has proved effective and does not require the use of additional and expensive equipment. The use of a low-toxic and nonexplosive carborane for boronization made this method of obtaining boron-carbon coatings to be a quick, widely available, and relatively cheap one. The coatings obtained were found to be highly resistant to chemical erosion—the erosion coefficients were (5–6) × 10−4 at/ion regardless of temperature. The electrical resistance of the coating was high, and, depending on the deposition conditions, varied in the range of 109–1011 Ω cm. The resistance of the coatings to the plasma impact was estimated using similar probes, which were examined after a certain number of working pulses. On all tokamaks, the coatings remained for several hundred pulses. The degradation of the coating correlates with the degradation of the plasma parameters [132,133,134,135,136].
The preparation of boron carbide through the pyrolysis of various carborane-containing polymers has been described [137,138,139,140]. The Ni-catalyzed polymerization of 1,2-bis(4-chloro-phenyl)-ortho-carborane leads to poly(phenylene-ortho-carborane). It was found that the heating of the polymer at 1000–1200 °C resulted in the crystallization of boron carbide, according to the X-ray powder diffraction studies [139]. The Ni-catalyzed polymerization of 1,7-bis(4-chlorophenyl)-meta-carborane produces poly(phenylene-meta-carborane), which can be used as a novel boron carbide precursor [140]. Due to its high ceramic yield, it can be used to prepare boron carbide ceramics with different shapes [141]. In particular, poly(phenylene-meta-carborane) was used to prepare the boron carbide hollow microsphere via slurry-coating and a method derived from a previous study. The poly(phenylene-meta-carborane)/polyacrylonitrile slurry was prepared and coated on a polyoxymethylene ball substrate. After air cross-linking, the substrate decomposition and heat-treatment at 1100 °C in argon atmosphere, hollow boron carbide microspheres with diameter of approximate 1.34 mm, and average shell thickness of 30 μm were obtained (Figure 6) [141].
The star-shaped pentagonal microcrystals of boron carbide with extremely low carbon content (~5%) were prepared through the thermobaric treatment of 1,7-bis(hydroxymethyl)-meta-carborane under high pressure of 7 GPa and temperature of 1370 K. The microcrystals exhibit a five-fold symmetry and grow in the shape of stars (Figure 7) [142,143]. The unusual shape of the pentagonal microcrystals makes them unique for developing novel micro-machines and semiconductor micro-devices [142].
Heating a mixture of ortho-carborane and adamantane (atomic ratio B:C = 5:95) at 8 GPa and 1700 °C results in the formation of boron-doped diamond microcrystals (2–2.5 at.% of boron), whereas only graphite was obtained from a mixture of adamantane and ortho-carborane at pressures lower than 7 GPa [144].

4. Boron Cluster Anions as Boron Source for Boron-Containing Materials

Coordination chemistry of transition metals with boron cluster anions is one of the most intensively studied fields of boron chemistry [145,146,147]. Research in this area is determined mainly by the fundamental components and concerns metal-boron cluster binding [148], positional isomerism [149,150], and secondary and interligand/inner-ligand interactions in complexes [151,152]. A series of new complex compounds that formed precursors and materials with desired properties were synthesized. Among them are precursors for the low-temperature synthesis of borides and related compounds [153], molecular switches based on a dimeric boron cluster [154], catalysts in the synthesis of organic compounds [155], complexes with luminescent properties [156], copper complexes as models for studying exchange processes and magnetic materials [157], as well as neutron-absorbing materials based on salts of boron cluster anions distributed in the silicate matrix.
Metal borides and related compounds provide ample opportunities for multivariate combination of metal–metal, metal–boron and boron–boron bonds in the resulting phases, thereby providing the possibility of directed changes in their physical, chemical and strength properties [157,158,159,160,161,162]. The most well-known methods of preparing metal borides include: (i) the reaction of metals and boron; (ii) the reduction of metal and boron from oxides when allowing to react with carbon or metals; (iii) the electrolytic reduction of metal and boron from their compounds; and (iv) the thermal dissociation of unstable compounds containing boron and metals. Actually, the processes used to prepare metal borides are often energy-consuming and time-consuming.
In the course of research carried out in our team, we have developed a method for obtaining binary borides during thermal reduction of transition metal compounds [MLx][An] (M = Co, Ni, An = [B10H10]2–, [B12H12]2– or [B20H18]2–) with ligands L that can be easily removed at elevated temperature (for example, L = H2O, NH3, DMF). In the compounds, organic ligands L are considered components, which play the role of organic fuel. Dimethylformamide is one of the most promising substances that can be used as a fuel [163]; its specific heat of combustion (29.652 MJ/kg) is much higher than, for example, that of urea (9.134 MJ/kg), which is often used in SCS processes. The energy capacity of the boron cluster anions themselves makes it possible to lower the boride synthesis temperature, which facilitates the process and reduces energy consumption.
First, we synthesized complexes [Co(DMF)6][B10H10] and [Co(DMSO)6][B10H10] (Figure 8), studied their thermooxidative properties in the temperature range 20–600 °C under argon [164], and determined the annealing temperature. When comparing the IR data of products of thermolysis performed at 600 °C, it was concluded that boride phases were prepared only for complex [Co(DMF)6][B10H10]. The final products were X-ray amorphous that did not allow us to determine the exactly composition of the final products.
When annealing structurally related compounds [Co(DMF)6][B12H12], [Co(DMF)6][B20H18] (Figure 9) and [Co(DMF)6][B10Cl10] in argon at 900 °C [165,166], we succeeded in detecting the CoB phase using X-ray powder diffraction [165]. It was found that for [Co(DMF)6][B12H12], the phases of BN and CoB where prepared in the 1:1 ratio; for [Co(DMF)6][B20H18], a higher CoB:BN ratio but low crystallinity were found; and for the cobalt(II) complex with the decachloro-closo-decaborate anion, only CoB was detected. The annealed samples were studied using IR spectroscopy and X-ray fluorescence (for the chloro-containing sample). The nanoparticular character of the decomposition products was shown using TEM.
Thermal reduction of complexes [CoLn][B10H10] (L = H2O, n = 6; N2H4, n = 3) with hydrazine and water molecules in argon at 650 and 900 °C [167,168] resulted in preparation of the dicobalt boride Co2B phase as well as orthorhombic and cubic modifications of boron nitride BN. For the aquacomplex, oxide-boride phases were detected. The annealed samples were studied using IR spectroscopy and X-ray powder diffraction. In addition, the samples show different magnetochemical behavior: the oxide–boride phase demonstrated a significant ferromagnetic contribution to the total magnetization of the sample, while the nitride–boride phase had a diamagnetic contribution.
As for structurally related nickel complexes [NiLn][B10H10] (L = DMF, H2O, n = 6; L = N2H4, n = 3) [169], their thermal reduction was studied in the temperature range 20–800 °C in air and in argon. The phases of Ni3C and Ni1 –xCx were detected using X-ray powder diffraction for annealed complex [Ni(DMF)6][B10H10]; the obtained data indicates that boride-carbide phases were not detected.
Gadolinium tetraboride GdB4 was found to form as an only-boride phase by heating a mixture of gadolinium hydride GdH~2 and gadolinium closo-decaborate Gd2[B10H10]3 as a boron source (the boron:metal ratio = 2) at 1400 °C under an argon atmosphere. In a similar way, cerium tetraboride CeB4 was prepared from CeH~2 and Ce2[B10H10]3 at 1100 °C. Using the boron:metal ratio = 6, gadolinium and cerium hexaborides MB6 (M = Gd, Ce) were prepared without the coexisting of the corresponding tetraborides at 1200–1400 °C and 1100 °C, respectively. A small number of inclusions (oxides, borates, etc.) can be completely removed using acid treatment with conc. HCI solution [170].
Crystalline ytterbium hexaboride YbB6 along with some amount of amorphous boron were prepared by heating ytterbium(II) closo-decaborate Yb[B10H10] in a quartz tube maintained at 10−5 Torr to a maximum of 1000 °C [171].
The thermal decomposition of copper(I) closo-decaborane Cu2[B10H10] at 800 °C was found to produce crystalline copper boride CuB24 and metal copper and amorphous boron [172].
Recently, the annealing of copper(II) complexes with hydrazine [CuII(N2H4)3][B10H10nH2O or ammonia [CuII(NH3)4][B10H10nH2O in argon at 900 °C was used to prepare a Cu@BN boron-containing copper composite [173]. The composition consists of a boron nitride matrix doped with cubic copper(0) nanoparticles with an average particle size of ~81 nm or ~52 nm, respectively.
Modern technology has a high demand for materials which can operate under extremal temperatures. Inorganic polymers attract attention because they offer some properties that are not found in organic materials, such as low-temperature flexibility, electrical conductivity, and nonflammability. The linear polysilicates obtained by the polycondensation of sodium metasilicate with silanol groups are the most widely studied among the non-organic polymers [174].
Prior to studying the distribution of salts of the boron cluster anions in the silicate matrix, the thermal and thermomechanical properties of starting salts (R3NH)2[B12H12] (R = Et, Bu) were examined as compared with (Et3NH)2[B10H10] [175]. The TGA and DSC data for (R3NH)2[B12H12] are similar; thermal destruction is observed at 260–450 °C, and the weakening of intermolecular contacts (softening) is observed before thermooxidative destruction. As for (Et3NH)2[B10H10], thermooxidative and thermal destructions occur simultaneously within a narrow temperature range of 260–320 °C, and the softening temperature lies within the range of intensive weight loss.
Furthermore, we studied the thermal behavior of triethylammonium closo-decaborate in a silicate matrix [176]. The interaction of sodium silicates of liquid glass (LG) with triethylammonium salts of boron cluster anions was studied in a wide range of component ratios. The compositions formed by addition of different amounts of (Et3NH)2[B10H10] (5, 15, 30, 40, 50, 60, and 74 wt%) into sodium liquid glass [176] were studied.
The dissolution of triethylammonium salts of boron cluster anions in sodium LG at room temperature is accompanied by the release of triethylamine, which completely stops when the temperature rises to 100 °C. The absence of a band of stretching vibrations of the NH groups of the triethylammonium cation in the region of 3100–3200 cm−1 indicates its complete replacement in the composition by Na+ ions. The retention of the formed sodium salts in the silicate matrix is carried out due to the formation of specific cation–anion contacts.
It was found that for compositions with closo-decaborate anion, the anion oxidation in air begins at 350 °C and is accompanied by a significant exothermic effect. IR spectroscopic analysis of the thermolysis products obtained in air at 350 and 600 °C showed the presence of the closo-decaborate anion in the samples [176,177]. A branched 3D system of multicenter bonds between BH-groups of the boron cluster and silanol groups via the water molecules can be assumed in the resulting inorganic polymer composition (Figure 10). The participation of the boron cluster in hydride–proton (dihydrogen) bonds is detected using IR spectroscopy because of the splitting of the band of stretching vibrations of the BH groups ν(BH) observed near 2500 cm−1, whereas the formation of hydrogen bonds between sylanol groups and water molecules can be assumed because of the presence of broadened band ν(OH) in the region 3600–3000 cm−1. This structure prevents the closo-decaborate anion from undergoing complete degradation, thus forming a surface protective layer which consists of borates and silicates, allowing preventing the bulky sample from oxygen diffusion and its further oxidation at high temperatures. The authors concluded that samples are stable up to 600 °C, which is attractive for fabricating boron-rich thermally stable coatings.
In the IR spectra of the compositions, the multiplet splitting of the band of stretching vibrations of BH bonds, which is characteristic of interactions of this kind, is clearly manifested. The thermal stability of individual salts of boron cluster anions is determined by the nature of the anion and cation of the starting compound. The thermal stability of the compositions also depends on the nature of the boron cluster anion. According to the TG and DSC data, the protective layer is formed when the temperature rises to 500 °C during the thermogravimetric analysis. It is worth noting that is that the heat treatment of the sample under these conditions is accompanied by a high exothermic effect, which can lead to the melting of the borosilicate components.
The possibility of using compositions based on the closo-decaborate anion as highly heat-resistant boron-enriched materials is evidenced by TMA data [178,179]. The samples are characterized by high heat resistance compared to the original components; they do not soften at temperatures ≥ 600 °C as high thermal and thermomechanical stability is probably ensured due to the formation of a “protective structure” on the surface of the samples during testing, which prevents the diffusion of atmospheric oxygen.
Structural features of boron cluster anions introduced into the compositions have a significant effect on the processes and thermomechanical properties of the compositions.
The formation of a protective layer in the compositions is also observed for the closo-dodecaborate anion [180,181]. It was found that heating the composition, in which the amount of the doped component is 60 wt%, leads to the formation of the composite and crystallization of sodium salt of the closo-dodecaborate anion on its surface at 200 °C, according to X-ray powder diffraction data [182]. The results of the study of the morphology of the obtained sample by scanning electron microscopy were compared with the results of the morphology of the sodium salt obtained from an aqueous solution. On the surface of the composite, there are needle-shaped nanosized particles with well-formed faces and sizes of 60–100 nm in width and up to 3 μm in length. In a sample of sodium salt obtained from an aqueous solution, only large blocks with a size of about 10–30 μm are present. In addition, when the initial mixture contained 60% of triethylammonium closo-dodecaborate at 450 °C, a high plasticization of the composition was noted, as evidenced by the TMA data [181]. The obtained properties may be important for the processing of composites and are probably due to the presence of about 6.6% triethylammonium substituted derivative of the closo-dodecaborate anion in the reaction, which is formed during the heat treatment of the composition.
The reaction of LG with the triethylammonium salt of the perchlorinated substituted derivative of the closo-decaborate anion proceeds similarly [179,180]. For LG/[B10Cl10]2– compositions containing the perchlorinated closo-decaborate anion up to 20 wt%, their plasticizing properties were determined. In the presence of small amounts of additives, associations formed between the silicate and polyhedral boron anions, which act as crosslinking agents. According to the TMA data, triethylammonium salt of the perchlorinated closo-decaborate anion does not soften up to a degradation temperature of 420 °C, whereas in salt (Et3NH)2[B10H10], this process occurs at 245 °C. Differences in the deformation stability are also retained in compositions containing equimolar amounts of boron cluster anions. Analyzing the obtained TMA results [179,180], it is obvious that the deformation stability of the system containing the perchlorinated anion is significantly higher compared to that of the decahydro-closo-decaborate anion. This fact indicates a more rigid structuring observed in the presence of the perchlorinated anion.
Compositions with a low content of boron cluster anions are of particular interest for studying the structural features of the associates formed. We suggested that the associates formed in the silicate matrix can be distributed as individual particles. which was determined using transmission electron microscopy (TEM). For samples containing the [B10H10]2– anion, the TEM image shows isolated elongated particles 12.5–47.5 nm in size. In turn, the shape of the particles for the composition with the perchlorinated anion is not so pronounced [180]. In the latter case, particles 5–40 nm in size form agglomerates distributed in a silicate matrix. Thus, it is obvious that the shape and nature of the distribution of associates formed in the silicate matrix directly depends on the nature of the boron cluster anion. As a result of these studies, we have patented a boron-containing neutron shielding material [183], which was obtained by the reaction between sodium silicate Na2O(SiO2)n in an aqueous solution of sodium hydroxide with trimethylammonium decahydro-closo-decaborate (Me3NH)2[B10H10]; the reaction solution was boiled until the trimethylamine formed as a result of the reaction of the sodium hydroxide solution with (Me3NH)2[B10H10] is completely removed, then dried by raising the temperature to 300 °C. Due to the numerous supramolecular contacts that appear in the glass structure, the destruction of the [B10H10]2– anion is not observed up to 600 °C. In addition, the boron content in the product is from 15 to 40 wt%, which provides a high ability of the material to capture thermal neutrons. The obtained neutron-shielding material can be used, in particular, in the encapsulation of radioactive waste, in the creation of protective shields.
Boron-containing compounds can be used as light components for the creation of metal matrix composites [184,185]. The metal composites containing copper and aluminum as matrices and salts Cs2[B10H10], [Me2NH2]2[B10H10], [Ph4P]2[B10H10], [Et3NH]2[B10Cl10], Cs2[B12H12], [Et3NH]2[B12H12] [149], and [Bu4N]2[B12H12] [150] were prepared and coated onto a steel surface. It was shown that the developed metal matrix composites with the boron cluster anion salts can be applied for coatings. A friction cladding method allows one to prepare high-quality coatings, providing a high adhesion of the coating to the metal substrate. No defects were found either in the mass of the coating or on the surface.

5. Conclusions

Here, we tried to summarize briefly the synthetic routes and wide application fields of boron-containing materials prepared from boron cluster anions and carboranes. The recent renaissance in chemistry of borohydrides and carboranes is associated with ever-new prospects for their practical use. We hope that the information collected in this article will significantly expand the understanding of the variability of the practical application of boron cluster anions and carboranes to obtain composites and materials based on them, and will provide a novel perspective on the ways to obtain composites with desired properties.

Author Contributions

Conceptualization, I.B.S. and E.A.M.; resources, S.E.N.; writing—original draft preparation, V.V.A.; writing—review and editing, I.B.S.; supervision, project administration, N.T.K. All authors have read and agreed to the published version of the manuscript.

Funding

The work was performed within the framework of the State Assignment of the Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences and the State Assignment of the Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences in the field of fundamental scientific research.

Data Availability Statement

No new data were created.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodger, E. Unconventional fuels. New Sci. 1957, 27–29. [Google Scholar]
  2. Martin, D.R. The development of borane fuels. J. Chem. Ed. 1959, 36, 208–214. [Google Scholar] [CrossRef]
  3. United States Government. 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.
  4. Sivaev, I.B. Nitrogen heterocyclic salts of polyhedral borane anions: From ionic liquids to energetic materials. Chem. Heterocycl. Comp. 2017, 53, 638–658. [Google Scholar] [CrossRef]
  5. Shan, Z.; Sheng, L.; Yang, R. Research progress of icosahedral polyhydroborate B12H122− anion compounds and their application in propellants and explosives. Chin. J. Explos. Propellants 2017, 40, 1–16. [Google Scholar] [CrossRef]
  6. Jiao, N.; Zhang, Y.; Liu, L.; Shreeve, J.M.; Zhang, S. Hypergolic fuels based on water-stable borohydride cluster anions with ultralow ignition delay times. J. Mater. Chem. A 2017, 5, 13341–13346. [Google Scholar] [CrossRef]
  7. 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]
  8. Sheng, L.; Shan, Z.; Guo, X.; Yang, R. Synthesis, characterization and thermal behavior of bis(dialkyl-5-aminotetrazolium) dodecadodecaborates. Chin. J. Org. Chem. 2018, 38, 2093–2100. [Google Scholar] [CrossRef]
  9. Zhang, Z.; Zhang, Y.; Li, Z.; Jiao, N.; Liu, L.; Zhang, S. B12H122–-Based metal (Cu2+, Ni2+, Zn2+) complexes as hypergolic fuels with superior hypergolicity. Eur. J. Inorg. Chem. 2018, 2018, 981–986. [Google Scholar] [CrossRef]
  10. Li, H.; Zhang, Y.; Liu, L.; Jiao, N.; Meng, X.; Zhang, S. Amino functionalized [B12H12]2– salts as hypergolic fuels. New J. Chem. 2018, 42, 3568–3573. [Google Scholar] [CrossRef]
  11. Derdziuk, J.; Malinowski, P.J.; Jaroń, T. Synthesis, structural characterization and thermal decomposition studies of (N2H5)2B12H12 and its solvates. Int. J. Hydrogen Energy 2019, 44, 27030–27038. [Google Scholar] [CrossRef]
  12. Sharon, P.; Afri, M.; Mitlin, S.; Gottlieb, L.; Schmerling, B.; Grinstein, D.; Welner, S.; Frimer, A.A. Preparation and characterization of bis(guanidinium) and bis(aminotetrazolium)dodecahydroborate salts: Green high energy nitrogen and boron rich compounds. Polyhedron 2019, 157, 71–89. [Google Scholar] [CrossRef]
  13. Li, S.; Pan, X.; Jiang, Y.; Chang, S.; Jin, X.; Yang, Y.; Huang, X.; Guo, Y. Ignition and combustion behaviors of high energetic polyhedral boron cluster. Propellants Explos. Pyrotech. 2019, 44, 1319–1326. [Google Scholar] [CrossRef]
  14. Wang, Q.-Y.; Wang, J.; Wang, S.; Wang, Z.-Y.; Cao, M.; He, C.-L.; Yang, J.-Q.; Zang, S.-Q.; Mak, T.C.W. o-Carborane-based and atomically precise metal clusters as hypergolic materials. J. Am. Chem. Soc. 2020, 142, 12010–12014. [Google Scholar] [CrossRef]
  15. Hawthorne, M.F. The role of chemistry in the development of boron neutron capture therapy of cancer. Angew. Chem. Int. Ed. 1993, 32, 950–984. [Google Scholar] [CrossRef]
  16. Soloway, A.H.; Tjarks, W.; Barnum, B.A.; Rong, F.-G.; Barth, R.F.; Codogni, I.M.; Wilson, J.G. The chemistry of neutron capture therapy. Chem. Rev. 1998, 98, 1515–1562. [Google Scholar] [CrossRef]
  17. Sivaev, I.B.; Bregadze, V.I. Polyhedral boranes for medical applications: Current status and perspectives. Eur. J. Inorg. Chem. 2009, 2009, 1433–1450. [Google Scholar] [CrossRef]
  18. Hu, K.; Yang, Z.; Zhang, L.; Xie, L.; Wang, L.; Xu, H.; Josephson, L.; Liang, S.H.; Zhang, M.-R. Boron agents for neutron capture therapy. Coord. Chem. Rev. 2020, 405, 213139. [Google Scholar] [CrossRef]
  19. Scholz, M.; Hey-Hawkins, E. Carbaboranes as pharmacophores: Properties, synthesis, and application strategies. Chem. Rev. 2011, 111, 7035–7062. [Google Scholar] [CrossRef]
  20. Fink, K.; Uchman, M. Boron cluster compounds as new chemical leads for antimicrobial therapy. Coord. Chem. Rev. 2021, 431, 213684. [Google Scholar] [CrossRef]
  21. Kugler, M.; Nekvinda, J.; Holub, J.; El Anwar, S.; Das, V.; Šícha, V.; Pospíšilová, K.; Fábry, M.; Král, V.; Brynda, J.; et al. Inhibitors of CA IX enzyme based on polyhedral boron compounds. ChemBioChem 2021, 22, 2741–2761. [Google Scholar] [CrossRef] [PubMed]
  22. Valliant, J.F.; Guenther, K.J.; King, A.S.; Morel, P.; Schaffer, P.; Sogbein, O.O.; Stephenson, K.A. The medicinal chemistry of carboranes. Coord. Chem. Rev. 2002, 232, 173–230. [Google Scholar] [CrossRef]
  23. Issa, F.; Kassiou, M.; Rendina, L.M. Boron in drug discovery: Carboranes as unique pharmacophores in biologically active compounds. Chem. Rev. 2011, 111, 5701–5722. [Google Scholar] [CrossRef]
  24. Zargham, E.O.; Mason, C.A.; Lee, M.W. The use of carboranes in cancer drug development. Int. J. Cancer Clin. Res. 2019, 6, 110. [Google Scholar] [CrossRef]
  25. Chen, Y.; Du, F.; Tang, L.; Xu, J.; Zhao, Y.; Wu, X.; Li, M.; Shen, J.; Wen, Q.; Cho, C.H.; et al. Carboranes as unique pharmacophores in antitumor medicinal chemistry. Mol. Ther. Oncolytics 2022, 24, 400–416. [Google Scholar] [CrossRef]
  26. Avdeeva, V.V.; Garaev, T.M.; Malinina, E.A.; Zhizhin, K.Y.; Kuznetsov, N.T. Physiologically active compounds based on membranotropic cage carriers—Derivatives of adamantane and polyhedral boron clusters (Review). Russ. J. Inorg. Chem. 2022, 67, 28–47. [Google Scholar] [CrossRef]
  27. Marforio, T.D.; Mattioli, E.J.; Zerbetto, F.; Calvaresi, M. Exploiting blood transport proteins as carborane supramolecular vehicles for boron neutron capture therapy. Nanomaterials 2023, 13, 1770. [Google Scholar] [CrossRef]
  28. Morin, J.F.; Sasaki, T.; Shirai, Y.; Guerrero, J.M.; Tour, J.M. Synthetic routes toward carborane-wheeled nanocars. J. Org. Chem. 2007, 72, 9481–9490. [Google Scholar] [CrossRef] [PubMed]
  29. Sasaki, T.; Guerrero, J.M.; Leonard, A.D.; Tour, J.M. Nanotrains and self-assembled two-dimensional arrays built from carboranes linked by hydrogen bonding of dipyridones. Nano Res. 2008, 1, 412–419. [Google Scholar] [CrossRef]
  30. Chiang, P.-T.; Mielke, J.; Godoy, J.; Guerrero, J.M.; Alemany, L.B.; Villagómez, C.J.; Saywell, A.; Grill, G.; Tour, J.M. Toward a light-driven motorized nanocar: Synthesis and initial imaging of single molecules. ACS Nano 2012, 6, 592–597. [Google Scholar] [CrossRef] [PubMed]
  31. Hosseini Lavasani, S.M.; Pishkenari, H.N.; Meghdari, A. How chassis structure and substrate crystalline direction affect the mobility of thermally driven p-carborane-wheeled nanocars. J. Phys. Chem. C 2019, 123, 4805–4824. [Google Scholar] [CrossRef]
  32. Hawthorne, M.F.; Zink, J.I.; Skelton, J.M.; Bayer, M.B.; Liu, C.; Livshits, E.; Baer, R.; Neuhauser, D. Electrical or photocontrol of the rotary motion of a metallacarborane. Science 2004, 303, 1849–1852. [Google Scholar] [CrossRef] [PubMed]
  33. Hawthorne, M.F.; Ramachandran, B.M.; Kennedy, R.D.; Knobler, C.B. Approaches to rotary molecular motors. Pure Appl. Chem. 2006, 78, 1299–1304. [Google Scholar] [CrossRef]
  34. Shlyakhtina, N.I.; Safronov, A.V.; Sevryugina, Y.V.; Jalisatgi, S.S.; Hawthorne, M.F. Synthesis, characterization, and preliminary fluorescence study of a mixed ligand bis(dicarbollyl)nickel complex bearing a tryptophan-BODIPY FRET couple. J. Organomet. Chem. 2015, 798, 234–244. [Google Scholar] [CrossRef]
  35. Sivaev, I.B. Ferrocene and transition metal bis(dicarbollides) as platform for design of rotatory molecular switches. Molecules 2017, 22, 2201. [Google Scholar] [CrossRef]
  36. Anufriev, S.A.; Timofeev, S.V.; Anisimov, A.A.; Suponitsky, K.Y.; Sivaev, I.B. Bis(dicarbollide) complexes of transition metals as a platform for molecular switches. Study of complexation of 8,8’-bis(methylsulfanyl) derivatives of cobalt and iron bis(dicarbollides). Molecules 2020, 25, 5745. [Google Scholar] [CrossRef]
  37. Bae, Y.S.; Farha, O.K.; Spokoyny, A.M.; Mirkin, C.A.; Hupp, J.T.; Snurr, R.Q. Carborane-based metal–organic frameworks as highly selective sorbents for CO2 over methane. Chem. Commun. 2008, 35, 4135–4137. [Google Scholar] [CrossRef]
  38. Spokoyny, A.M.; Farha, O.K.; Mulfort, K.L.; Hupp, J.T.; Mirkin, C.A. Porosity tuning of carborane-based metal–organic frameworks (MOFs) via coordination chemistry and ligand design. Inorg. Chim. Acta 2010, 364, 266–271. [Google Scholar] [CrossRef]
  39. Kennedy, R.D.; Krungleviciute, V.; Clingerman, D.J.; Mondloch, J.E.; Peng, Y.; Wilmer, C.E.; Sarjeant, A.A.; Snurr, R.Q.; Hupp, J.T.; Yildirim, T.; et al. Carborane-based metal-organic framework with high methane and hydrogen storage capacities. Chem. Mater. 2013, 25, 3539–3543. [Google Scholar] [CrossRef]
  40. Kennedy, R.D.; Clingerman, D.J.; Morris, W.; Wilmer, C.E.; Sarjeant, A.A.; Stern, C.L.; O’Keeffe, M.; Snurr, R.Q.; Hupp, J.T.; Farha, O.K.; et al. Metallacarborane-based metal–organic framework with a complex topology. Cryst. Growth Des. 2014, 14, 1324–1330. [Google Scholar] [CrossRef]
  41. Clingerman, D.J.; Morris, W.; Mondloch, J.E.; Kennedy, R.D.; Sarjeant, A.A.; Stern, C.; Hupp, J.T.; Farha, O.K.; Mirkin, C.A. Stabilization of a highly porous metal–organic framework utilizing a carborane-based linker. Chem. Commun. 2015, 51, 6521–6523. [Google Scholar] [CrossRef]
  42. Boldog, I.; Bereciartua, P.J.; Bulánek, R.; Kučeráková, M.; Tomandlová, M.; Dušek, M.; Macháček, J.; De Vos, D.; Baše, T. 10-Vertex closo-carborane: A unique ligand platform for porous coordination polymers. CrystEngComm 2016, 18, 2036–2040. [Google Scholar] [CrossRef]
  43. Boldog, I.; Dušek, M.; Jelínek, T.; Švec, P.; de Ramos, O.F.S.; Růžička, A.; Bulánek, R. Porous 10- and 12-vertex (bi)-p-dicarba-closo-boranedicarboxylates of cobalt and their gas adsorptive properties. Microporous Mesoporous Mater. 2018, 271, 284–294. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Hu, J.; Krishna, R.; Wang, L.; Yang, L.; Cui, X.; Duttwyler, S.; Xing, H. Rational design of microporous MOFs with anionic boron cluster functionality and cooperative dihydrogen binding sites for highly selective capture of acetylene. Angew. Chem. Int. Ed. 2020, 59, 17664–17669. [Google Scholar] [CrossRef]
  45. Sun, W.; Hu, J.; Duttwyler, S.; Wang, L.; Krishna, R.; Zhang, Y. Highly selective gas separation by two isostructural boron cluster pillared MOFs. Sep. Purif. Technol. 2022, 283, 120220. [Google Scholar] [CrossRef]
  46. Li, Z.; Li, X.-B.; Light, M.E.; Carrillo, A.E.; Arauzo, A.; Valvidares, M.; Roscini, C.; Teixidor, F.; Viñas, C.; Gándara, F.; et al. A metal-organic framework incorporating eight different size rare-earth metal elements: Toward multifunctionality À La Carte. Adv. Funct. Mater. 2023, 33, 2307369. [Google Scholar] [CrossRef]
  47. Sivaev, I.B. Decaborane: From Alfred Stock and rocket fuel projects to nowadays. Molecules 2023, 28, 6287. [Google Scholar] [CrossRef]
  48. Grimes, R.N. Carboranes, 3rd ed.; Academic Press: London, UK, 2016. [Google Scholar] [CrossRef]
  49. Sivaev, I.B.; Prikaznov, A.V.; Naoufal, D. Fifty years of the closo-decaborate anion chemistry. Collect. Czechoslov. Chem. Commun. 2010, 75, 1149–1199. [Google Scholar] [CrossRef]
  50. Zhizhin, K.Y.; Zhdanov, A.P.; Kuznetsov, N.T. Derivatives of closo-decaborate anion [B10H10]2− with exo-polyhedral substituents. Russ. J. Inorg. Chem. 2010, 55, 2089–2127. [Google Scholar] [CrossRef]
  51. 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]
  52. Hawthorne, M.F.; Shelly, K.; Li, F. The versatile chemistry of the [B20H18]2− ions: Novel reactions and structural motifs. Chem. Commun. 2002, 6, 547–554. [Google Scholar] [CrossRef]
  53. Avdeeva, V.V.; Malinina, E.A.; Zhizhin, K.Y.; Bernhardt, E.; Kuznetsov, N.T. Structural diversity of dimer clusters based on the octadecahydro-eicosaborate anion. J. Struct. Chem. 2019, 60, 692–712. [Google Scholar] [CrossRef]
  54. Thompson, A.G. MOCVD technology for semiconductors. Mater. Lett. 1997, 30, 255–263. [Google Scholar] [CrossRef]
  55. Jones, A.C. Recent developments in the MOCVD of electronic materials. Adv. Mater. Opt. Electron. 2000, 10, 91–92. [Google Scholar] [CrossRef]
  56. Chow, L.A. Equipment and manufacturability issues in chemical vapor deposition processes. In Handbook of Thin Film Deposition, 4th ed.; Seshan, K., Schepis, D., Eds.; William Andrew: Oxford, UK, 2018; pp. 269–316. [Google Scholar] [CrossRef]
  57. Global Metal Organic Chemical Vapor Deposition Market: Analysis by Application, by Category, by Region Size and Trends with Impact of COVID-19 and Forecast up to 2026. Available online: https://www.researchandmarkets.com/report/mocvd (accessed on 1 June 2023).
  58. Roush, G. The Toxicology of the Boranes. J. Occup. Med. 1959, 1, 46–52. Available online: https://www.jstor.org/stable/44999044 (accessed on 1 June 2023). [CrossRef]
  59. Naeger, L.L.; Leibman, K.C. Mechanisms of decaborane toxicity. Toxicol. Appl. Pharmacol. 1972, 22, 517–527. [Google Scholar] [CrossRef]
  60. 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]
  61. Kamimura, K.; Yoshimura, T.; Nagaoka, T.; Nakao, M.; Onuma, Y.; Makimura, M. Preparation and thermoelectric property of boron thin film. J. Solid State Chem. 2000, 154, 153–156. [Google Scholar] [CrossRef]
  62. Natsir, M.; Sagara, A.; Motojima, O. Reduction of hydrogen content in boron film by controlling glow discharge conditions. J. Nucl. Mater. 1995, 220–222, 865–868. [Google Scholar] [CrossRef]
  63. 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]
  64. 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]
  65. 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]
  66. 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]
  67. 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]
  68. 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]
  69. 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, 22, 3214–3215. [Google Scholar] [CrossRef] [PubMed]
  70. 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]
  71. 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]
  72. 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]
  73. 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]
  74. 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]
  75. 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]
  76. 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]
  77. 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]
  78. Su, K.; Sneddon, L.G. A polymer precursor route to metal borides. Chem. Mater. 1993, 5, 1659–1668. [Google Scholar] [CrossRef]
  79. 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]
  80. Borchardt, L.; Kockrick, E.; Wollmann, P.; Kaskel, S.; Guron, M.M.; Sneddon, L.G.; Geiger, D. Ordered mesoporous boron carbide based materials via precursor nanocasting. Chem. Mater. 2010, 22, 4660–4668. [Google Scholar] [CrossRef]
  81. 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]
  82. 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]
  83. 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 main chain. Chem. Commun. 2014, 50, 4585–4587. [Google Scholar] [CrossRef] [PubMed]
  84. 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. Des. 2016, 109, 408–414. [Google Scholar] [CrossRef]
  85. Welna, D.T.; Bender, J.D.; Wei, X.; Sneddon, L.G.; Allcock, H.R. Preparation of boron-carbide/carbon nanofibers from a poly(norbornenyldecaborane) single-source precursor via electrostatic spinning. Adv. Mater. 2005, 17, 859–862. [Google Scholar] [CrossRef]
  86. 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] [CrossRef]
  87. 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]
  88. Wang, J.; Gou, Y.; Zhang, Q.; Jian, K.; Chen, Z.; Wang, H. Linear organodecaborane block copolymer as a single-sourceprecursor for porous boron carbide ceramics. J. Eur. Ceram. Soc. 2017, 37, 1937–1943. [Google Scholar] [CrossRef]
  89. 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]
  90. 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]
  91. Mazurowski, J.; Lee, S.; Ramseyer, G.; Dowben, P.A. Characterization of boron carbide films formed by PECVD. MRS Online Proc. Libr. 1992, 242, 637–642. [Google Scholar] [CrossRef]
  92. Lee, S.; Mazurowski, J.; Ramseyer, G.; Dowben, P.A. Characterization of boron carbide thin films fabricated by plasma enhanced chemical vapor deposition from boranes. J. Appl. Phys. 1992, 72, 4925–4933. [Google Scholar] [CrossRef]
  93. Lee, S.; Mazurowski, J.; O’Brien, W.L.; Dong, Q.Y.; Jia, J.J.; Callcott, T.A.; Tan, Y.; Miyano, K.E.; Ederer, D.L.; Mueller, D.R.; et al. The structural homogeneity of boron carbide thin films fabricated using plasma enhanced chemical vapor deposition from B5H9+CH4. J. Appl. Phys. 1993, 74, 6919–6924. [Google Scholar] [CrossRef]
  94. 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 1 June 2023).
  95. Rees, W.S.; Seyferth, D. High-yield synthesis of B4C/BN ceramic materials by pyrolysis of polymeric Lewis base adducts of decaborane(14). J. Am. Ceram. Soc. 1988, 71, C194–C196. [Google Scholar] [CrossRef]
  96. Rees, W.S.; Seyferth, D. Preparation, characterization, and pyrolysis of decaborane (14)—Based polymers: B4C/BN and BN procedures. In A Collection of Papers Presented at the 13th Annual Conference on Composites and Advanced Ceramic Materials: Ceramic Engineering and Science Proceedings; The American Ceramic Society: Westerville, OH, USA, 1989; Volume 10, pp. 837–845. [Google Scholar] [CrossRef]
  97. Yogo, T.; Naka, S. Synthesis of boron nitride from triammoniadecaborane and hydrazine under pressure. J. Mater. Sci. 1990, 25, 374–378. [Google Scholar] [CrossRef]
  98. Yogo, T.; Naka, S.; Iwahara, H. Synthesis of cubic boron nitride with boron nitride powder formed from triammoniadecaborane. J. Mater. Sci. 1991, 26, 3758–3762. [Google Scholar] [CrossRef]
  99. Robertson, B.W.; Adenwalla, S.; Harken, A.; Welsch, P.; Brand, J.I.; Dowben, P.A.; Claassen, J.P. A class of boron-rich solid-state neutron detectors. Appl. Phys. Lett. 2002, 80, 3644–3646. [Google Scholar] [CrossRef]
  100. Harken, A.D.; Day, E.E.; Robertson, B.W.; Adenwalla, S. Boron-rich semiconducting boron carbide neutron detector. Jpn. J. Appl. Phys. 2005, 44, 444–445. [Google Scholar] [CrossRef]
  101. Day, E.; Diaz, M.J.; Adenwalla, S. Effect of bias on neutron detection in thin semiconducting boron carbide films. J. Phys. D 2006, 39, 2920–2924. [Google Scholar] [CrossRef]
  102. Caruso, A.N.; Dowben, P.A.; Balkir, S.; Schemm, N.; Osberg, K.; Fairchild, R.W.; Flores, O.B.; Balaz, S.; Harken, A.D.; Robertson, B.W.; et al. The all boron carbide diode neutron detector: Comparison with theory. Mater. Sci. Eng. B 2006, 135, 129–133. [Google Scholar] [CrossRef]
  103. Hong, N.; Crow, L.; Adenwalla, S. Time-of-flight neutron detection using PECVD grown boron carbide diode detector. Nucl. Instrum. Methods Phys. Res. Sect. A 2013, 708, 19–23. [Google Scholar] [CrossRef]
  104. Peterson, G.; Su, Q.; Wang, Y.; Dowben, P.A.; Nastasi, M. Improved p–n heterojunction device performance induced by irradiation in amorphous boron carbide films. Mater. Sci. Eng. B 2015, 202, 25–30. [Google Scholar] [CrossRef]
  105. Nastasi, M.; Peterson, G.; Su, Q.; Wang, Y.; Ianno, N.J.; Benker, N.; Echeverría, E.; Yost, J.A.; Kelber, A.J.; Dong, B.; et al. Electrical and structural characterization of neutron irradiated amorphous boron carbide/silicon p-n heterojunctions. Nucl. Inst Methods Phys. Res. B 2018, 432, 48–54. [Google Scholar] [CrossRef]
  106. Byun, D.; Spady, B.R.; Ianno, N.J.; Dowben, P.A. Comparison of different chemical vapor deposition methodologies for fabrication of hetero junction boron-carbide diodes. Nanostruct. Mater. 1995, 5, 465–471. [Google Scholar] [CrossRef]
  107. Zhang, D.; Mcilroy, D.N.; O’Brien, W.L.; De Stasio, G. The chemical and morphological properties of boron-carbon alloys grown by plasma-enhanced chemical vapour deposition. J. Mater. Sci. 1998, 33, 4911–4915. [Google Scholar] [CrossRef]
  108. Nordell, B.J.; Karki, S.; Nguyen, T.D.; Rulis, P.; Caruso, A.N.; Purohit, S.S.; Li, H.; King, S.W.; Dutta, D.; Gidley, D.; et al. The influence of hydrogen on the chemical, mechanical, optical/electronic, and electrical transport properties of amorphous hydrogenated boron carbide. J. Appl. Phys. 2015, 118, 035703. [Google Scholar] [CrossRef]
  109. Bute, A.; Jagannath; Kar, R.; Chopade, S.S.; Desai, S.S.; Deo, M.N.; Rao, P.; Chand, N.; Kumar, S.; Singh, K.; et al. Effect of self-bias on the elemental composition and neutron absorption of boron carbide films deposited by RF plasma enhanced CVD. Mater. Chem. Phys. 2016, 182, 62–71. [Google Scholar] [CrossRef]
  110. Nordell, B.J.; Keck, C.L.; Nguyen, T.D.; Caruso, A.N.; Purohit, S.S.; Lanford, W.A.; Dutta, D.; Gidley, D.; Henry, P.; King, S.W.; et al. Tuning the properties of a complex disordered material: Full factorial investigation of PECVD-grown amorphous hydrogenated boron carbide. Mater. Chem. Phys. 2016, 173, 268–284. [Google Scholar] [CrossRef]
  111. Bute, A.; Jena, S.; Bhattacharya, D.; Kumar, S.; Chand, N.; Keskar, N.; Sinha, S. Composition dependent microstructure and optical properties of boron carbide (BxC) thin films deposited by radio frequency-plasma enhanced chemical vapour deposition technique. Mater. Res. Bull. 2019, 109, 175–182. [Google Scholar] [CrossRef]
  112. Tu, R.; Hu, X.; Li, J.; Yang, M.; Li, Q.; Shi, J.; Li, H.; Ohmori, H.; Goto, T.; Zhang, S. Fabrication of (a-nc) boron carbide thin films via chemical vapor deposition using ortho-carborane. J. Asian Ceram. Soc. 2020, 8, 327–335. [Google Scholar] [CrossRef]
  113. Bute, A.; Jena, S.; Kedia, S.; Udupa, D.V.; Singh, K.; Bhattacharya, D.; Modi, M.H.; Chand, N.; Sinha, S. Boron carbide thin films deposited by RF-PECVD and PLD technique: A comparative study based on structure, optical properties, and residual stress. Mater. Chem. Phys. 2021, 258, 123860. [Google Scholar] [CrossRef]
  114. Bute, A.; Jena, S.; Sharma, R.K.; Jagannath, U.D.V.; Maiti, N. Linear and non-linear optical properties of boron carbide thin films. Appl. Surf. Sci. 2023, 608, 155101. [Google Scholar] [CrossRef]
  115. Caruso, A.N.; Billa, R.B.; Balaz, S.; Brand, J.I.; Dowben, P.A. The heteroisomeric diode. J. Phys. Condens. Matter. 2004, 16, L139–L146. [Google Scholar] [CrossRef]
  116. Caruso, A.N.; Balaz, S.; Xu, B.; Dowben, P.A.; McMullen-Gunn, A.S.; Brand, J.I.; Losovyj, Y.B.; McIlroy, D.N. Surface photovoltage effects on the isomeric semiconductors of boron-carbide. Appl. Phys. Lett. 2004, 84, 1302–1304. [Google Scholar] [CrossRef]
  117. Echeverría, E.; Dong, B.; Liu, A.; Wilson, E.R.; Peterson, G.; Nastasi, M.; Dowben, P.A.; Kelber, J.A. Strong binding at the gold (Au) boron carbide interface. Surf. Coat. Technol. 2017, 314, 51–54. [Google Scholar] [CrossRef]
  118. Caruso, A.N. The physics of solid-state neutron detector materials and geometries. J. Phys. Condens. Matter 2010, 22, 443201. [Google Scholar] [CrossRef]
  119. Peterson, G.G.; Echeverria, E.; Dong, B.; Silva, J.P.; Wilson, E.R.; Kelber, J.A.; Nastasi, M.; Dowben, P.A. Increased drift carrier lifetime in semiconducting boron carbides deposited by plasma enhanced chemical vapor deposition from carboranes and benzene. J. Vac. Sci. Technol. A 2017, 35, 03E101. [Google Scholar] [CrossRef]
  120. Oyelade, A.; Yost, A.J.; Benker, N.; Dong, B.; Knight, S.; Schubert, M.; Dowben, P.A.; Kelber, J.A. Composition-dependent charge transport in boron carbides alloyed with aromatics: Plasma enhanced chemical vapor deposition aniline/ortho-carborane films. Langmuir 2018, 34, 12007–12016. [Google Scholar] [CrossRef]
  121. Dong, B.; James, R.; Kelber, J.A. PECVD of boron carbide/aromatic composite films: Precursor stability and resonance stabilization energy. Surf. Coat. Technol. 2016, 290, 94–99. [Google Scholar] [CrossRef]
  122. Dong, B.; Oyelade, A.; Nandagopal, N.; Kelber, J.A. Aromatic-doped boron carbide films formed by PECVD of metacarborane and aniline or pyridine: Chemical and electronic structure. Surf. Coat. Technol. 2017, 314, 45–50. [Google Scholar] [CrossRef]
  123. Echeverría, E.; James, R.; Chiluwal, U.; Pasquale, F.L.; Colón Santana, J.A.; Gapfizi, R.; Tae, J.-D.; Driver, M.S.; Enders, A.; Kelber, J.A.; et al. Novel semiconducting boron carbide/pyridine polymers for neutron detection at zero bias. Appl. Phys. A 2015, 118, 113–118. [Google Scholar] [CrossRef]
  124. Echeverria, E.; Dong, B.; Peterson, G.; Silva, J.P.; Wilson, E.R.; Driver, M.S.; Jun, Y.-S.; Stucky, G.D.; Knight, S.; Hofmann, T.; et al. Semiconducting boron carbides with better charge extraction through the addition of pyridine moieties. J. Phys. D 2016, 49, 355302. [Google Scholar] [CrossRef]
  125. Oyelade, A.; Osonkie, A.; Yost, A.J.; Benker, N.; Dowben, P.A.; Kelber, J.A. Optical, electronic and visible-range photo-electronic properties of boron carbide-indole films. J. Phys. D 2020, 53, 355101. [Google Scholar] [CrossRef]
  126. Hwang, S.-D.; Remmes, N.B.; Dowben, P.A.; McIlroy, D.N. Nickel doping of boron carbide grown by plasma enhanced chemical vapor deposition. J. Vac. Sci. Technol. B 1996, 14, 2957–2960. [Google Scholar] [CrossRef]
  127. Hwang, S.-D.; Remmes, N.; Dowben, P.A.; McIlroy, D.N. Nickel doping of boron-carbon alloy films and corresponding Fermi level shifts. J. Vac. Sci. Technol. B 1997, 15, 854–859. [Google Scholar] [CrossRef]
  128. Hwang, S.-D.; Yang, K.; Dowben, P.A.; Ahmad, A.A.; Ianno, N.J.; Li, J.Z.; Lin, J.Y.; Jiang, H.X.; McIlroy, D.N. Fabrication of n-type nickel doped B5C1+δ homojunction and heterojunction diodes. Appl. Phys. Lett. 1997, 70, 1028–1030. [Google Scholar] [CrossRef]
  129. Carlson, L.; LaGraffe, D.; Balaz, S.; Ignatov, A.; Losovyj, Y.B.; Choi, J.; Dowben, P.A.; Brand, J.I. Doping of boron carbides with cobalt, using cobaltocene. Appl. Phys. A 2007, 89, 195–201. [Google Scholar] [CrossRef]
  130. Ignatov, A.Y.; Losovyj, Y.B.; Carlson, L.; LaGraffe, D.; Brand, J.I.; Dowben, P.A. Pairwise cobalt doping of boron carbides with cobaltocene. J. Appl. Phys. 2007, 102, 083520. [Google Scholar] [CrossRef]
  131. Dowben, P.A.; Kizilkaya, O.; Liu, J.; Montag, B.; Nelson, K.; Sabirianov, I.; Brand, J.I. 3d transition metal doping of semiconducting boron carbides. Mater. Lett. 2009, 63, 72–74. [Google Scholar] [CrossRef]
  132. Sharapov, V.M.; Mirnov, S.V.; Grashin, S.A.; Lebedev, S.V.; Kovan, I.A.; Krasilnikov, A.V.; Krupin, V.A.; Levin, L.S.; Romannikov, A.N.; Zakharov, A.P. Boronization of Russian tokamaks from carborane precursors. J. Nucl. Mater. 1995, 220–222, 730–735. [Google Scholar] [CrossRef]
  133. Hong, S.H.; Lee, K.S.; Kim, K.M.; Kim, H.T.; Kim, G.P.; Sun, J.H.; Woo, H.J.; Park, J.M.; Kim, W.C.; Kim, H.K.; et al. KSTAR team. First boronization in KSTAR. Fusion Eng. Des. 2010, 85, 946–949. [Google Scholar] [CrossRef]
  134. Hong, S.-H.; Lee, K.-S.; Kim, K.-P.; Kim, K.-M.; Kim, H.-T.; Sun, J.-H.; Woo, H.-J.; Park, J.-M.; Park, F.-K.; Kim, W.-C.; et al. First boronization in KSTAR: Experiences on carborane. J. Nucl. Mater. 2011, 415, S1050–S1053. [Google Scholar] [CrossRef]
  135. Ko, J.; Den Hartog, D.J.; Goetz, J.A.; Weix, P.J.; Limbach, S.T. First gaseous boronization during pulsed discharge cleaning. J. Nucl. Mater. 2013, 432, 146–151. [Google Scholar] [CrossRef]
  136. Sharapov, V.M. Discharge chamber plasma-chemical conditioning in magnetic confinement fusion devices (Review). Phys. Atom. Nucl. 2021, 84, 1266–1271. [Google Scholar] [CrossRef]
  137. Zheng, H.; Thorne, K.; Mackenzie, J.D.; Yang, X.; Hawthorne, M.F. Boron carbide-based ceramics via polymer route synthesis. MRS Online Proc. Libr. 1991, 249, 15–23. [Google Scholar] [CrossRef]
  138. Johnson, S.E.; Yang, X.; Hawthorne, M.F.; Thorne, K.J.; Zheng, H.; Mackenzie, J.D. Alkynyl substituted carboranes as precursors to boron carbide thin films, fibers and composites. MRS Online Proc. Libr. 1992, 271, 833–838. [Google Scholar] [CrossRef]
  139. Cheng, S.; Yuan, K.; Wang, X.; Han, J.; Jian, X.; Wang, J. Poly(phenylene-carborane) for boron-carbide/carbon ceramic precursor synthesized via nickel catalysis. Polymer 2017, 115, 224–231. [Google Scholar] [CrossRef]
  140. Yan, D.; Chen, J.; Zhang, Y.; Gou, Y. Synthesis and characterization of a carborane-containing precursor for B4C ceramics. Sci. Discov. 2021, 9, 128–132. [Google Scholar] [CrossRef]
  141. Yan, D.; Chen, J.; Zhang, Y.; Gou, Y. Preparation of novel carborane-containing boron carbide precursor and its derived ceramic hollow microsphere. Ceram. Int. 2022, 48, 18392–18400. [Google Scholar] [CrossRef]
  142. Filonenko, V.P.; Zinin, P.V.; Zibrov, I.P.; Anokhin, A.S.; Kukueva, E.V.; Lyapin, S.G.; Fominski, V.Y. Synthesis of star-shaped boron carbide micro-crystallites under high pressure and high temperatures. Crystals 2018, 8, 448. [Google Scholar] [CrossRef]
  143. Pavlov, I.S.; Ivanova, A.G.; Filonenko, V.P.; Zibrov, I.P.; Voloshin, A.E.; Zinin, P.V.; Vasiliev, A.L. The rhombic hexecontahedronboron carbide microcrystals—Crystal structure analysis. Scr. Mater. 2023, 222, 115023. [Google Scholar] [CrossRef]
  144. Bagramov, R.H.; Filonenko, V.P.; Zibrov, I.P.; Skryleva, E.A.; Nikolaev, A.V.; Pasternak, D.G.; Vlasov, I.I. Highly boron-doped graphite and diamond synthesized from adamantane and ortho-carborane under high pressure. Materialia 2022, 21, 101274. [Google Scholar] [CrossRef]
  145. Avdeeva, V.V.; Malinina, E.A.; Kuznetsov, N.T. Boron cluster anions and their derivatives in complexation reactions. Coord. Chem. Rev. 2022, 469, 214636. [Google Scholar] [CrossRef]
  146. Matveev, E.Y.; Avdeeva, V.V.; Zhizhin, K.Y.; Malinina, E.A.; Kuznetsov, N.T. Effect of nature of substituents on coordination properties of mono- and disubstituted derivatives of boron cluster anions [BnHn]2– (n = 10, 12) and carboranes with exo-polyhedral B-X Bonds (X = N, O, S, Hal). Inorganics 2022, 10, 238. [Google Scholar] [CrossRef]
  147. Paskevicius, M.; Hansen, B.R.S.; Jørgensen, M.; Richter, B.; Jensenet, T.R. Multifunctionality of silver closo-boranes. Nat. Commun. 2017, 8, 15136. [Google Scholar] [CrossRef] [PubMed]
  148. 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]
  149. Avdeeva, V.V.; Malinina, E.A.; Kuznetsov, N.T. Isomerism in salts and complexes with boron cluster anions [B10H10]2– and [B20H18]2–. Russ. J. Inorg. Chem. 2020, 65, 335–358. [Google Scholar] [CrossRef]
  150. Avdeeva, V.V.; Malinina, E.A.; Kuznetsov, N.T. Isomerism in complexes with the decahydro-closo-decaborate anion. Polyhedron 2016, 105, 205–221. [Google Scholar] [CrossRef]
  151. Kravchenko, E.A.; Gippius, A.A.; Tkachev, A.V.; Zhurenko, S.V.; Golubev, A.V.; Kubasov, A.S.; Selivanov, N.A.; Buzanov, G.A.; Bykov, A.Y.; Zhizhin, K.Y.; et al. Non-covalent Cl⋯X interactions in several silver compounds based on [B12Cl12]2− clusters: 35Cl NQR and X-ray diffraction. Inorg. Chim. Acta 2023, 544, 121231. [Google Scholar] [CrossRef]
  152. Avdeeva, V.V.; Kravchenko, E.A.; Gippius, A.A.; Vologzhanina, A.V.; Malinina, E.A.; Zhurenko, S.V.; Buzanov, G.A.; Kuznetsov, N.T. Decachloro-closo-decaborate anion in copper(II) complexation reactions with N-donor ligands: 35Cl NQR and X-ray studies. Polyhedron 2017, 127, 238–247. [Google Scholar] [CrossRef]
  153. Albert, B.; Hofmann, K. Metal Borides: Versatile Structures and Properties. In Handbook of Solid State Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2017. [Google Scholar] [CrossRef]
  154. Avdeeva, V.V.; Buzin, M.I.; Dmitrienko, A.O.; Dorovatovskii, P.V.; Malinina, E.A.; Kuznetsov, N.T.; Voronova, E.D.; Zubavichus, Y.V.; Vologzhanina, A.V. Solid-state reactions of eicosaborate [B20H18]2− salts and complexes. Chem. Eur. J. 2017, 23, 16819–16828. [Google Scholar] [CrossRef]
  155. Eleazer, B.J.; Smith, M.D.; Peryshkov, D.V. Reaction of a ruthenium B-carboranyl hydride complex and BH3(SMe2): Selective formation of a pincer-supported metallaborane LRu(B3H8). Tetrahedron 2019, 75, 1471–1474. [Google Scholar] [CrossRef]
  156. Korolenko, S.E.; Zhuravlev, K.P.; Tsaryuk, V.I.; Kubasov, A.S.; Avdeeva, V.V.; Malinina, E.A.; Burlov, A.S.; Divaeva, L.N.; Zhizhin, K.Y.; Kuznetsov, N.T. Crystal structures, luminescence, and DFT study of mixed-ligand Zn(II) and Cd(II) complexes with phenyl-containing benzimidazole derivatives with linker CN or NN group. J. Lumin. 2021, 237, 118156. [Google Scholar] [CrossRef]
  157. Malinina, E.A.; Vologzhanina, A.V.; Avdeeva, V.V.; Goeva, L.V.; Efimov, N.N.; Ugolkova, E.A.; Minin, V.V.; Kuznetsov, N.T. Structures, magnetic properties, and EPR studies of tetranuclear copper(II) complexes [Cu4(OH)4L4]4+ (L = bpa, bipy) stabilized by anions containing decahydro-closo-decaborate anion. Polyhedron 2020, 183, 114540. [Google Scholar] [CrossRef]
  158. Chung, H.-Y.; Weinberger, M.B.; Levine, J.B.; Kavner, A.; Yang, J.-M.; Tolbert, S.H.; Kaner, R.B. Synthesis of ultra-incompressible superhard rhenium diboride at ambient pressure. Science 2007, 316, 436–439. [Google Scholar] [CrossRef]
  159. Latini, A.; Rau, J.V.; Teghil, R.; Generosi, A.; Albertini, V.R. Superhard properties of rhodium and iridium boride films. ACS Appl. Mater. Interfaces 2010, 2, 581–587. [Google Scholar] [CrossRef]
  160. Xie, M.; Mohammadi, R.; Turner, C.L.; Kaner, R.B.; Kavner, A.; Tolbert, S.H. Lattice stress states of superhard tungsten tetraboride from radial X-ray diffraction under nonhydrostatic compression. Phys. Rev. B Condens. Matter Mater. Phys. 2014, 90, 104104. [Google Scholar] [CrossRef]
  161. Bazhin, P.; Chizhikov, A.; Bazhina, A.; Konstantinov, A.; Avdeeva, V. Titanium-titanium boride matrix composites prepared in-situ under conditions combining combustion processes and high-temperature shear deformation. Mater. Sci. Eng. A 2023, 874, 145093. [Google Scholar] [CrossRef]
  162. Prokopets, A.D.; Bazhin, P.M.; Konstantinov, A.S.; Antipov, M.S.; Avdeeva, V.V. Structural features of layered composite material TiB2/TiAl/Ti6Al4 obtained by unrestricted SHS-compression. Mater. Lett. 2021, 300, 130165. [Google Scholar] [CrossRef]
  163. Petrichko, M.I.; Karavaev, I.A.; Savinkina, E.V.; Grigoriev, M.S.; Buzanov, G.A.; Retivov, V.M. Rare-earth nitrate complexes with dimethylformamide. Russ. J. Inorg. Chem. 2023, 68, 415–423. [Google Scholar] [CrossRef]
  164. Avdeeva, V.V.; Polyakova, I.N.; Vologzhanina, A.V.; Goeva, L.V.; Buzanov, G.A.; Generalova, N.B.; Malinina, E.A.; Zhizhin, K.Y.; Kuznetsov, N.T. [Co(solv)6][B10H10] (solv = DMF and DMSO) for low-temperature synthesis of borides. Russ. J. Inorg. Chem. 2016, 61, 1125–1134. [Google Scholar] [CrossRef]
  165. Malinina, E.A.; Myshletsov, I.I.; Buzanov, G.A.; Kubasov, A.S.; Kozerozhets, I.V.; Goeva, L.V.; Nikiforova, S.E.; Avdeeva, V.V.; Zhizhin, K.Y.; Kuznetsov, N.T. A New approach to the synthesis of nanocrystalline cobalt boride in the course of the thermal decomposition of cobalt complexes [Co(DMF)6]2+ with boron cluster anions. Molecules 2023, 28, 453. [Google Scholar] [CrossRef] [PubMed]
  166. Avdeeva, V.V.; Vologzhanina, A.V.; Ugolkova, E.A.; Minin, V.V.; Malinina, E.A.; Kuznetsov, N.T. Synthesis and structures of compounds [ML6][B10Cl10] (M = Co, Ni; L = CH3CN, DMF, DMSO) as precursors for synthesis of cobalt(II) and nickel(II) complexes with organic L ligands. J. Solid State Chem. 2021, 296, 121989. [Google Scholar] [CrossRef]
  167. Malinina, E.A.; Goeva, L.V.; Buzanov, G.A.; Avdeeva, V.V.; Efimov, N.N.; Kozerozhets, I.V.; Kuznetsov, N.T. Synthesis and physicochemical properties of binary cobalt(II) borides. Thermal reduction of precursor complexes [CoLn][B10H10] (L = H2O, n = 6; N2H4, n = 3). Russ. J. Inorg. Chem. 2019, 64, 1325–1334. [Google Scholar] [CrossRef]
  168. Malinina, E.A.; Goeva, L.V.; Buzanov, G.A.; Avdeeva, V.V.; Efimov, N.N.; Kuznetsov, N.T. A new method for synthesis of binary borides with desired properties. Dokl. Chem. 2019, 487, 180–183. [Google Scholar] [CrossRef]
  169. Malinina, E.A.; Goeva, L.V.; Buzanov, G.A.; Retivov, V.M.; Avdeeva, V.V.; Kuznetsov, N.T. Synthesis and thermal reduction of complexes [NiLn][B10H10] (L = DMF, H2O, n = 6; L = N2H4, n = 3): Formation of solid solutions Ni3C1 –xBx. Russ. J. Inorg. Chem. 2020, 65, 126–132. [Google Scholar] [CrossRef]
  170. Itoh, H.; Tsuzuki, Y.; Yogo, T.; Naka, S. Synthesis of cerium and gadolinium borides using boron cage compounds as a boron source. Mater. Res. Bull. 1987, 22, 1259–1266. [Google Scholar] [CrossRef]
  171. White III, J.P.; Shore, S.G. Complexes of divalent lanthanides (ytterbium(II), europium(II), and samarium(II)) with decaborates. Inorg. Chem. 1992, 31, 2756–2761. [Google Scholar] [CrossRef]
  172. Volkov, V.V.; Yur’ev, G.S.; Solomatina, L.Y.; Voronina, G.S.; Myakishev, K.G. Thermal changes in decahydro-closo-decaborate(2-)copper(I) Cu2B10H10 and the nature of copper borides. Bull. Acad. Sci. USSR Div. Chem. Sci. 1990, 39, 435–438. [Google Scholar] [CrossRef]
  173. Malinina, E.A.; Myshletsov, I.I.; Buzanov, G.A.; Kozerozhets, I.V.; Simonenko, N.P.; Simonenko, T.L.; Nikiforova, S.E.; Avdeeva, V.V.; Zhizhin, K.Y.; Kuznetsov, N.T. Physicochemical fundamentals of the synthesis of a Cu@BN composite consisting of nanosized copper enclosed in a boron nitride matrix. Inorganics 2023, 11, 345. [Google Scholar] [CrossRef]
  174. Skachkova, V.K.; Grachev, A.V.; Shaulov, A.Y.; Berlin, A.A. Inorganic polymers using sodium silicate liquid glass. Features of silicate polycondensation. Russ. J. Phys. Chem. B 2019, 13, 849–852. [Google Scholar] [CrossRef]
  175. Skachkova, V.K.; Goeva, L.V.; Grachev, A.V.; Avdeeva, V.V.; Malinina, E.A.; Shaulov, A.Y.; Berlin, A.A.; Kuznetsov, N.T. Thermal and thermomechanical properties of trialkylammonium dodecahydro-closo-dodecaborates (R3NH)2[B12H12] (R = Et, Bu). Russ. J. Inorg. Chem. 2017, 62, 84–89. [Google Scholar] [CrossRef]
  176. Goeva, L.V.; Skachkova, V.K.; Avdeeva, V.V.; Malinina, E.A.; Grachev, A.V.; Shaulov, A.Y.; Berlin, A.A.; Kuznetsov, N.T. Interactions of sodium liquid glass with triethylammonium decahydro-closo-decaborate (Et3NH)2B10H10. Russ. J. Inorg. Chem. 2014, 59, 107–110. [Google Scholar] [CrossRef]
  177. Skachkova, V.K.; Goeva, L.V.; Grachev, A.V.; Avdeeva, V.V.; Malinina, E.A.; Shaulov, A.Y.; Berlin, A.A.; Kuznetsov, N.T. Thermal oxidation of the decahydro-closo-decaborate anion B10H102− in a silicate matrix. Inorg. Mater. 2015, 51, 498–502. [Google Scholar] [CrossRef]
  178. Skachkova, V.K.; Malinina, E.A.; Goeva, L.V.; Grachev, A.V.; Avdeeva, V.V.; Kozerozhets, I.V.; Shaulov, A.Y.; Berlin, A.A.; Kuznetsov, N.T. Morphology of supramolecular structures based on sodium silicates and [B10R10]2– (R = H, Cl) boron cluster anions. Inorg. Mater. 2021, 57, 1173–1177. [Google Scholar] [CrossRef]
  179. Malinina, E.; Skachkova, V.; Goeva, L.; Grachev, A.; Lyubimova, G.; Avdeeva, V.; Kozerozhets, I.; Shaulov, A.; Berlin, A.; Kuznetsov, N. Thermomechanical properties of compositions based on polysilicates modified with boron cluster anions or SiO2 nanoparticles. Bol. Soc. Esp. Ceram. Vidr. 2020, 59, 201–208. [Google Scholar] [CrossRef]
  180. Skachkova, V.K.; Malinina, E.A.; Goeva, L.V.; Grachev, A.V.; Avdeeva, V.V.; Shaulov, A.Y.; Berlin, A.A.; Kuznetsov, N.T. Polycondensation of water glass sodium silicates in the presence of [BnXn]2– (n = 10, 12; X = H, Cl) boron cluster anions. Inorg. Mater. 2020, 56, 657–661. [Google Scholar] [CrossRef]
  181. Skachkova, V.K.; Goeva, L.V.; Grachev, A.V.; Kochneva, I.K.; Malinina, E.A.; Shaulov, A.Y.; Berlin, A.A.; Kuznetsov, N.T. Composites based on triethylammonium dodecahydro-closo-dodecaborate ((Et 3NH)2[B12H12]) and sodium silicate water glass. Inorg. Mater. 2017, 53, 207–211. [Google Scholar] [CrossRef]
  182. Malinina, E.A.; Skachkova, V.K.; Kozerozhets, I.V.; Avdeeva, V.V.; Goeva, L.V.; Buzanov, G.A.; Shaulov, A.Y.; Berlin, A.A.; Kuznetsov, N.T. Formation of nanoscale sodium dodecahydro-closo-dodecaborate Na2[B12H12] on the surface of a silicate matrix. Dokl. Chem. 2019, 484, 1–4. [Google Scholar] [CrossRef]
  183. Goeva, L.V.; Malinina, E.A.; Avdeeva, V.V.; Kuznetsov, N.T.; Skachkova, V.K.; Shaulov, A.Y.; Grachev, A.V.; Berlin, A.A. Boron-Containing Neutron Shielding Material. RF Patent RU2550156C1, 10 May 2016. [Google Scholar]
  184. Popov, V.A.; Zhizhin, K.Y.; Malinina, E.A.; Ketsko, V.A.; Kuznetsov, N.T. A possibility of using mechanical alloying for developing metal matrix composites with light-weight reinforcements. J. Alloys Compd. 2007, 434–435, 451–454. [Google Scholar] [CrossRef]
  185. Popov, V.A.; Zhizhin, K.Y.; Kuznetsov, N.T.; Staudhammer, K.P.; Retivov, V.M. Investigation of the possibility of application of boron clusters in composite materials with metal matrix. Adv. Mater. Res. 2009, 59, 96–100. [Google Scholar] [CrossRef]
Figure 1. Idealized structures of decaborane(14) B10H14 (a), ortho-carborane 1,2-C2B10H12 (b), meta-carborane 1,7-C2B10H12 (c), decahydro-closo-decaborate anion [B10H10]2− (d), and octadecahydro-conjucto-eicosoborate anion [trans-B20H18]2− (e).
Figure 1. Idealized structures of decaborane(14) B10H14 (a), ortho-carborane 1,2-C2B10H12 (b), meta-carborane 1,7-C2B10H12 (c), decahydro-closo-decaborate anion [B10H10]2− (d), and octadecahydro-conjucto-eicosoborate anion [trans-B20H18]2− (e).
Materials 16 06099 g001
Figure 2. Solid-state structure of [μ-6,6′-(CH2)6-(B10H13)2]. Hydrogen atoms of organic substituents are omitted for clarity.
Figure 2. Solid-state structure of [μ-6,6′-(CH2)6-(B10H13)2]. Hydrogen atoms of organic substituents are omitted for clarity.
Materials 16 06099 g002
Figure 3. SEM images of aligned boron carbide nanofibers obtained upon pyrolysis of [μ-6,6′-(CH2)6-(B10H13)2] at 1025 °C. Reprinted with permission from Ref. [79]. Copyright (2000) the American Chemical Society.
Figure 3. SEM images of aligned boron carbide nanofibers obtained upon pyrolysis of [μ-6,6′-(CH2)6-(B10H13)2] at 1025 °C. Reprinted with permission from Ref. [79]. Copyright (2000) the American Chemical Society.
Materials 16 06099 g003
Figure 4. SEM images of ceramic fibers at different scale (A,B) derived from [-6-B10H12-(ethylenediamine)]n- by pyrolysis at 1000 °C under argon. Reprinted with permission from Ref. [87]. Copyright (1988) the American Ceramic Society.
Figure 4. SEM images of ceramic fibers at different scale (A,B) derived from [-6-B10H12-(ethylenediamine)]n- by pyrolysis at 1000 °C under argon. Reprinted with permission from Ref. [87]. Copyright (1988) the American Ceramic Society.
Materials 16 06099 g004
Figure 5. Surface and cross-section SEM images of boron carbide film prepared at Tdep = 900 °C, Ptot = 100 Pa (a,e), Tdep = 900 °C, Ptot = 50,000 Pa (b,f), Tdep = 1100 °C, Ptot = 5000 Pa (c,g), Tdep = 1200 °C, Ptot = 50,000 Pa (d,h). Reprinted from Ref. [112].
Figure 5. Surface and cross-section SEM images of boron carbide film prepared at Tdep = 900 °C, Ptot = 100 Pa (a,e), Tdep = 900 °C, Ptot = 50,000 Pa (b,f), Tdep = 1100 °C, Ptot = 5000 Pa (c,g), Tdep = 1200 °C, Ptot = 50,000 Pa (d,h). Reprinted from Ref. [112].
Materials 16 06099 g005
Figure 6. Hollow boron carbide microspheres prepared from poly(phenylene-meta-carborane)/polyacrylonitrile at different magnitude (af). Reprinted with permission from Ref. [141]. Copyright (2022) Elsevier.
Figure 6. Hollow boron carbide microspheres prepared from poly(phenylene-meta-carborane)/polyacrylonitrile at different magnitude (af). Reprinted with permission from Ref. [141]. Copyright (2022) Elsevier.
Materials 16 06099 g006
Figure 7. SEM images of star-shaped pentagonal boron carbide microcrystals prepared by thermobaric treatment of 1,7-bis(hydroxymethyl)-meta-carborane at 7 GPa and 1370 K. Reprinted from Ref. [142].
Figure 7. SEM images of star-shaped pentagonal boron carbide microcrystals prepared by thermobaric treatment of 1,7-bis(hydroxymethyl)-meta-carborane at 7 GPa and 1370 K. Reprinted from Ref. [142].
Materials 16 06099 g007
Figure 8. Structures of [Co(DMSO)6][B10H10].
Figure 8. Structures of [Co(DMSO)6][B10H10].
Materials 16 06099 g008
Figure 9. Structure of [Co(DMF)6][B20H18].
Figure 9. Structure of [Co(DMF)6][B20H18].
Materials 16 06099 g009
Figure 10. Multicenter interactions between (Et3NH)2[B10H10] and components of sodium liquid glass.
Figure 10. Multicenter interactions between (Et3NH)2[B10H10] and components of sodium liquid glass.
Materials 16 06099 g010
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

Avdeeva, V.V.; Nikiforova, S.E.; Malinina, E.A.; Sivaev, I.B.; Kuznetsov, N.T. Composites and Materials Prepared from Boron Cluster Anions and Carboranes. Materials 2023, 16, 6099. https://doi.org/10.3390/ma16186099

AMA Style

Avdeeva VV, Nikiforova SE, Malinina EA, Sivaev IB, Kuznetsov NT. Composites and Materials Prepared from Boron Cluster Anions and Carboranes. Materials. 2023; 16(18):6099. https://doi.org/10.3390/ma16186099

Chicago/Turabian Style

Avdeeva, Varvara V., Svetlana E. Nikiforova, Elena A. Malinina, Igor B. Sivaev, and Nikolay T. Kuznetsov. 2023. "Composites and Materials Prepared from Boron Cluster Anions and Carboranes" Materials 16, no. 18: 6099. https://doi.org/10.3390/ma16186099

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

Article Metrics

Back to TopTop