Next Article in Journal
A Structural Study of 0.06LiNbO3-0.94K0.5Na0.5NbO3 from Neutron Total Scattering Analysis
Next Article in Special Issue
New Radical-Cation Salts Based on the TMTTF and TMTSF Donors with Iron and Chromium Bis(Dicarbollide) Complexes: Synthesis, Structure, Properties
Previous Article in Journal
Mechanical Properties of CaO–Al2O3–SiO2 Glass-Ceramics Precipitating Hexagonal CaAl2Si2O8 Crystals
Previous Article in Special Issue
Carboranes as Lewis Acids: Tetrel Bonding in CB11H11 Carbonium Ylide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

1,12-Diiodo-Ortho-Carborane: A Classic Textbook Example of the Dihalogen Bond †

by
Kyrill Yu. Suponitsky
1,*,
Alexei A. Anisimov
1,2,
Sergey A. Anufriev
1,
Igor B. Sivaev
1,* and
Vladimir I. Bregadze
1
1
A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilov Str., 119991 Moscow, Russia
2
Higher Chemical College at the Russian Academy of Sciences, D.I. Mendeleev Russian Chemical Technological University, 9 Miusskaya Sq., 125047 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Dedicated to Professor Alan J. Welch in occasion of his retirement at Herriot-Watt University and in recognition of his outstanding contribution in the carborane chemistry.
Crystals 2021, 11(4), 396; https://doi.org/10.3390/cryst11040396
Submission received: 15 March 2021 / Revised: 29 March 2021 / Accepted: 6 April 2021 / Published: 8 April 2021
(This article belongs to the Special Issue Carborane: Dedicated to the Work of Professor Alan Welch)

Abstract

:
The crystal structure of 1,12-diiodo-ortho-carborane 1,12-I2-1,2-C2B10H10 was determined by single crystal X-ray diffraction. In contrary to earlier studied 1,12-dibromo analogue 1,12-Br2-1,2- C2B10H10, its crystal packing is governed by the presence of the intermolecular I⋯I dihalogen bonds between the iodine atom attached to the carbon atom (acceptor) and the iodine atom attached to the antipodal boron atom (donor) of the carborane cage. The observed dihalogen bonds belong to the II type and are characterized by classical parameters: shortened I⋯I distance of 3.5687(9) Å, C–I⋯I angle of 172.61(11)° and B–I⋯I angle of 92.98(12)°.

1. Introduction

Carboranes [CB11H12] and C2B10H12, in which one or two vertices in boron icosahedron are replaced by a carbon unit, are a fascinating family of compounds with exceptional chemical and thermal robustness, unique geometry, rigidity, and synthetic versatility [1]. Selective chemical substitution of hydrogen atoms at carbon or boron atoms in these clusters allows for their use as rigid, three dimensional scaffolds upon which to construct new materials, such as liquid crystals [2,3,4,5], nonlinear optical materials [6,7,8,9], carborane-based anticrowns [10], and even in drug design [11,12]. As rigid molecules of fixed length, carboranes can be used as building blocks (“molecular tinkertoys”) [13,14,15,16,17,18] for supramolecular assemblies, such as porous coordination polymers or metal–organic frameworks (MOFs) [19,20,21,22]. Another type of supramolecular structures with the participation of carboranes is based on the acidity of their CH groups, which demonstrate a high potential for hydrogen bonding. Indeed, intermolecular C–H⋯O and C–H⋯N hydrogen bonding, including bifurcated interactions, features in much of the supramolecular chemistry of carboranes [23,24]. Functionalization of carboranes with different substituents including halogen atoms opens an opportunity to the formation diverse noncovalent interactions [25,26]. Thus, the intermolecular C–H⋯X–B hydrogen bonds were found to stabilize crystal structures of fluoro- [27], bromo- [28] and iodo- [29,30,31,32,33,34] derivatives of ortho-carborane. Alternatively, hydrogen atoms bonded to the carbon atoms can also be replaced by halogen atoms. It should be noted that when substituted at a carbon atom, carborane acts as an electron-withdrawing group with respect to a substituent, while when substituted at boron atoms it plays the role of an electron-releasing group. The further the location of a substituent is from carbon atoms, the higher the electron-releasing ability is the carborane cage [35].
Halogen bonds are one of the strongest noncovalent intermolecular interactions, and are formed between the σ-hole of a halogen atom and nucleophile [36,37,38,39]. In the case of halogen bonds in which both atoms are halogens, the σ-hole is activated by electron acceptor substitution of a halogen while donor substituents are necessary to increase ability of lone pair donation of the second halogen atom. Therefore, in the case when halogen atoms are simultaneously introduced to the carborane carbon atom and the boron atom is antipodal to it, this makes the formation of intermolecular halogen bonds possible, where the halogen atom attached to the carbon atom plays the role of an acceptor, and the halogen atom bonded to boron acts as a donor. Thus, the B,C-dihalogen-substituted carboranes represent a unique class of small molecules, in the crystals of which the formation of intermolecular dihalogen bonds could be possible without the participation of the second component. In particular, one might expect a formation of the intermolecular dihalogen bonds for 1,12-dihalo-ortho-carboranes 1,12-X2-1,2-C2B10H10. However, the recent study of the crystal structure of 1,12-Br2-1,2-C2B10H10 showed that in this case, instead of the formation of the intermolecular C-Br⋯Br-B dihalogen bonds, the formation of the C–H⋯Br–B hydrogen and C-Br⋯H-B halogen bonds occurs [28]. On the other hand, the σ-hole size, which is the determining factor in the formation of a halogen bond, depends on both the electronic effect of the substituent and the electronegativity of the halogen atom [40,41,42]. This prompted us to study intermolecular interactions in an analogous diiodine derivative 1,12-I2-1,2-C2B10H10 using single crystal X-ray diffraction and quantum chemical calculations.

2. Results and Discussion

Despite the fact that the syntheses of the C-iodo derivatives of ortho-carborane were first reported more than 50 years ago [43,44], they were on the periphery of mainstream carborane chemistry developments, and were not even well characterized [34,45]. In this way, they radically differ from the B-iodo derivatives of ortho-carborane, which have found active use in the synthesis of B-alkyl and aryl derivatives by means of Pd-catalyzed cross-coupling reactions [46,47,48,49,50,51,52]. Therefore, the synthesis of the C-iodo derivatives of ortho-carborane is not an easy task.
In this respect, the synthesis and characterization of the C-halogen derivatives of the carba-closo-dodecaborate anion [1-X-1-CB11H11] (X = F, Cl, Br, I) are described much better [53]. Moreover, the preparation of its 1,12-diiodo derivative [1,12-I2-1-CB11H10], containing iodine atoms in opposite positions of the boron backbone, has recently been described [54]. However, in the case of anionic carboranes, it is rather difficult to find a cation that, on the one hand, will not form additional non-covalent bonds with the anion and, on the other hand, will be small enough not to hinder the formation of intermolecular dihalogen bonds between the anions. In addition, unlike the C-substituted ortho-carborane, the C-substituted carba-closo-dodecaborate anion has no or negligible electron-withdrawing effects [55].
An attempt to prepare 1,12-I2-1,2-C2B10H10 by the reaction of the lithium derivative of 9-iodo-ortho-carborane with iodine in 1,2-dimethoxyethane resulted in the expected formation of a mixture of 1,9-I2-1,2-C2B10H10 and 1,12-I2-1,2-C2B10H10 derivatives (1:1). However, in contrast to the similar dibromo derivatives 1,9-Br2-1,2-C2B10H10 and 1,12-Br2-1,2-C2B10H10 [56], we failed to separate this mixture. Nevertheless, we managed to obtain the desired 1,12-diiodo derivative as a by-product of the cross-coupling reaction of 9-iodo-ortho-carborane with phenylmagnesium bromide. Another by-product of this reaction was 1-iodo-ortho-carborane, which we also obtained by direct reaction of the lithium derivative of ortho-carborane with iodine. Notably, formation of similar products of iodine migration in the process of cross-coupling of B-iodo carboranes with Grignard reagents was noted earlier [57].
As mentioned in the Introduction, no Br⋯Br halogen bond was observed in the crystal structure of 1,12-dibromo-ortho-carborane 1,12-Br2-1,2-C2B10H10 [28]. It should be noted that similarities and differences between bonding preferences of the bromine atom in comparison to iodine atom, on the one hand, and chlorine atom, on the other hand, was the subject of extensive studies [58,59,60,61]. Based on comparison of the crystal packing of 1-Ph-2-X-ortho-carboranes (X = F, Cl, Br, I), it was shown that both Br and I form Hal⋯π interactions, while neither Cl or F participate in such interactions [58]. Study of N-(2-halo-2,2-dinitroethyl)pyrrolidine- 2,5-diones (Hal = F, Cl, Br) [59] has revealed that both Cl and Br participate in halogen bonding, but bromine interacts with the carbonyl oxygen atom (the strongest donor site), while chlorine prefers to connect to much weaker donors, namely, oxygen atoms of the nitro group. Based on the above, it becomes unclear a priori which packing motif should be expected in the crystal of 1,12-diiodo-ortho-carborane.
Single crystals of 1,12-I2-1,2-C2B10H10 suitable for X-ray study were obtained in the form of thin plates by slow evaporation of chloroform solution. An asymmetric unit cell of 1,12-I2-1,2-C2B10H10 contains one molecule (Figure 1). The I1–C1 bond length (2.121(2) Å) is slightly longer than average X-ray value for I–C (aromatic) bonds (2.095 Å [62]) and is significantly shorter than the B12–I12 bond (2.179(2) Å).
The main packing motif of the crystal structure of 1,12-I2-1,2-C2B10H10 is represented by infinite chains along the b direction formed by the C–I⋯I–B dihalogen bonds of II type [37,63] (the I(12)⋯I(1’) distance is 3.5687(9) Å, the B(12)-I(12)⋯I(1’) and I(12)⋯I(1’)-C(1’) angles are 92.98(12) and 172.61(11)°, respectively) (Figure 2), which is very different from 1,12-Br2-1,2-C2B10H10 studied earlier.
In order to find out the reason of observed differences as well as peculiarities of the crystal packing of 1,12-I2-1,2-C2B10H10, we used energetic analysis of intermolecular contacts, that is frequently invoked for crystal packing study [64,65,66]. We calculated halogen bonded dimer for both compound 1,12-I2-1,2-C2B10H10 and similar dimer in which iodine atoms are replaced with bromines. The results are presented in Figure 3 and Table 1. The calculated dimer of 1,12-I2-1,2-C2B10H10 is characterized by the structure similar to that found experimentally. The I⋯I distance is somewhat shorter, while C–I⋯I and B–I⋯I angles and B–H⋯I distances are close to experimentally observed values. Topological analysis of calculated electron density for 1,12-I2-1,2-C2B10H10 dimer has revealed additional stabilization of the dimeric structure with the B–H⋯I hydride–halogen bonds that was not evident from the consideration of bare X-ray data. From Table 1, it can be seen that energy of the I⋯I contact is sizably higher than that of the B–H⋯I contacts; therefore, the I⋯I dihalogen bond can be considered as the structure-forming interaction in the crystal of 1,12-I2-1,2-C2B10H10. In contrary, optimized geometry of dimeric 1,12-dibromo-ortho- carborane appeared to be quite different. The C–Br⋯Br angle significantly deviates from 180°. As a consequence, energy of the Br⋯Br interactions is relatively small and becomes comparable to the B–H⋯Br interactions which are also formed between two molecules in the dimer. It means that the Br⋯Br interactions are no more structure-forming ones. These results are in qualitative agreement with a previous experiment [26]; according to which, no Br⋯Br halogen bond is observed in the crystal of 1,12-Br2-1,2-C2B10H10.
The above results have demonstrated computational ability for, at least, qualitative explanation and prediction of the main crystal packing motif for dihalogen derivatives of ortho-carborane. Based on that, we made an attempt to predict the possibility of halogen bond formation in 1,3-I2-1,2-C2B10H10 and 1,9-I2-1,2-C2B10H10. Those isomers were chosen because they can be experimentally obtained from the available 3- and 9-iodo-ortho-carboranes, while synthesis of other possible isomers is troublesome. In Figure 4 and Table 2, the results of calculation of dimers of the 1,3- and 1,9-isomers are presented.
One can see that 1,9-I2-1,2-C2B10H10 demonstrates the same system of close contacts and nearly the same energetic properties of halogen-bonded dimer of 1,12-I2-1,2-C2B10H10. The energy of a halogen bond is only 0.2 kcal/mol less; B9-I9⋯I1’ and I9⋯I1’-C1’ only slightly deviate from 90 and 180°, respectively, while energies of B–H⋯I interactions are only 0.1 kcal/mol higher. It allows consideration of the I⋯I halogen bond as a predominant interaction in the potential crystal structure of 1,9-I2-1,2-C2B10H10.
When looking at halogen-bonded dimers built up of 1,3-I2-1,2-C2B10H10, one can observe a clear trend of weakening of the I⋯I halogen bonds and simultaneous strengthening of the B–H⋯I interactions and increases in their roles in stabilizing dimeric structures upon approaching the iodine substituent from its position at B12 to B3. During movement of the iodine atom from B12 to B3, quantitative changes due to the weakening of the I⋯I halogen bonds can be transformed to qualitative changes, which can result in the disappearance of the I⋯I halogen bonds from the crystal structure of 1,3-I2-1,2-C2B10H10.
In conclusion, 1,12-diiodo-ortho-carborane was obtained and its crystal structure was determined by X-ray diffraction, which revealed the existence of the I⋯I halogen bond in its crystal structure, in contrast to 1,12-dibromo-ortho-carborane. Based on quantum chemical calculation, we have determined preferences of the type of noncovalent interactions in 1,12-diiodo- and 1,12-dibromo-ortho-carboranes which appeared to be in agreement with experimental findings. Based on our results, we can predict the formation of the I⋯I halogen bonds in 1,9-diiodo-ortho-carborane, while our results cannot provide solid support for the formation of such bonds in the 1,3-isomer. This question is still open and can be answered experimentally. Synthesis and crystal growth of 1,9- and 1,3-diiodo-ortho- carboranes is in progress in our group.

3. Materials and Methods

3.1. General

Compounds 9-iodo-ortho-carborane and bis(triphenylphosphine)palladium(II) dichloride were prepared according to the literature procedures [67,68]. Solvents 1,2-dimethoxyethane and diethyl ether were dried using standard procedures [69]. Phenyl iodide was distilled at boiling point. All other chemical reagents were purchased from Sigma Aldrich, Acros Organics and ABCR and used without purification. All reactions were carried out in an argon atmosphere. The reaction progress was monitored by thin-layer chromatography (Merck F254 silica gel on aluminum plates) and visualized using 0.5% PdCl2 in 1% HCl in aq. MeOH (1:10). Acros Organics silica gel (0.060–0.200 mm) was used for column chromatography. The NMR spectra at 400.1 MHz (1H), 128.4 MHz (11B) and 100.0 MHz (13C) were recorded with a Varian Inova 400 spectrometer. The residual signal of the NMR solvent relative to Me4Si was taken as an internal reference for 1H and 13C NMR spectra; 11B NMR spectra were referenced using BF3.Et2O as an external standard.

3.2. Cross-Coupling of 9-Iodo-Ortho-Carborane with PhMgBr

Phenyl iodide (0.70 mL, 1275 mg, 6.25 mmol) was added to a mixture of magnesium turnings (228 mg, 9.38 mmol) in fresh distilled diethyl ether (25 mL). The resulting mixture was heated under reflux for 1 h. Then, 9-iodo-ortho-carborane (675 mg, 2.50 mmol) in fresh distilled diethyl ether (25 mL) was added, and the reaction was stirred at room temperature for another 1 h. Then, copper(I) iodide (25 mg, 0.13 mmol, catalytic amount) with [(Ph3P)2PdCl2] (83 mg, 0.13 mmol, catalytic amount) were added. The reaction was heated under reflux for 16 h and 6% HCl in water (50 mL) was added. The organic layer was separated; the water layer was washed with diethyl ether (3 × 50 mL). The organic phases were combined, dried over Na2SO4, and concentrated under reduced pressure. The crude product was purified by column chromatography on silica using a mixture of chloroform and petroleum ether (1:3, v/v) to give, along with the expected 9-phenyl-ortho-carborane, pale-yellow solids of 1-iodo-ortho-carborane (15 mg, yield 2%) and 1,12-diiodo-ortho- carborane (20 mg, yield 2%) as side products.
Compound 1-I-1,2-C2B10H11: 1H NMR (CDCl3, ppm): δ 3.78 (1H, br s, CHcarb), 3.7–0.7 (10H, br m, BH). 11B NMR (CDCl3, ppm): δ −0.7 (1B, d, J = 151 Hz), −4.0 (1B, d, J = 154 Hz), −7.8 (4B, d, J = 125 Hz), −9.0 (2B, d, J = 111 Hz), −11.7 (1B, d, J = 167 Hz).
Compound 1,12-I2-1,2-C2B10H10: 1H NMR (CDCl3, ppm): δ 3.86 (1H, br s, CHcarb), 4.1–0.6 (9H, br m, BH). 11B NMR (CDCl3, ppm): δ 0.9 (1B, d, J = 156 Hz), −5.8 (2B, d, J = 160 Hz), −7.8 (2B, d, J = 173 Hz), −8.7 (2B, d, J = 183 Hz), −10.7 (2B, d, J = 168 Hz), −16.7 (1B, s).

3.3. General Synthetic Procedure of C-Iodination of Ortho-Carborane and Its B-I Derivatives

The 2.25 M BuLi in hexanes was added to a mixture of carborane in fresh distilled 1,2-dimethoxyethane (10 mL). The mixture was stirred for 1 h at room temperature and I2 was added by one portion. The reaction was stirred at room temperature overnight and Na2S2O3·5H2O (1000 g, 4.03 mmol) in water (10 mL) and diethyl ether (15 mL) were added. The organic layer was separated; the water layer was washed with diethyl ether (25 mL). The organic phases were combined, dried over Na2SO4, and concentrated under reduced pressure. The crude product was purified by column chromatography on silica using diethyl ether as the eluent to give the corresponding C–I derivative.
Compound 1,2-I2-1,2-C2B10H10: 2.25 M BuLi in hexanes (1.00 mL, 2.25 mmol), ortho-carborane (144 mg, 1.00 mmol) and I2 (635 mg, 2.50 mmol) were used; a yellow crystalline solid was obtained (381 mg, yield 96%). 1H NMR (CDCl3, ppm): δ 3.9–0.8 (10H, br m, BH). 11B NMR (CDCl3, ppm): δ −2.5 (2B, d, J = 154 Hz), −4.5 (2B, d, J = 183 Hz), −6.8 (6B, d, J = 183 Hz).
Compounds 1,9- and 1,12-I2-1,2-C2B10H10: 2.25 M BuLi in hexanes (0.49 mL, 1.10 mmol), 9-iodo-ortho-carborane (135 mg, 0.50 mmol) and I2 (305 mg, 1.20 mmol) were used; a pale-grey crystalline solid was obtained (80 mg, yield 20%). 1H NMR (CDCl3, ppm): δ 4.07 (1H, br s, CHcarb, 1,9-isomer), 3.87 (1H, br s, CHcarb, 1,12-isomer), 3.9–0.6 (20H, br m, BH, 1,9 + 1,12-isomers).

3.4. Synthesis of 1-Iodo-Ortho-Carborane

The 2.25 M BuLi in hexanes (0.40 mL, 0.90 mmol) was added to a mixture of ortho-carborane (144 mg, 1.00 mmol) in fresh distilled 1,2-dimethoxyethane (10 mL). The mixture was stirred for 1 h at room temperature, and I2 (381 mg, 1.50 mmol) was added by one portion. The reaction was stirred at room temperature overnight and Na2S2O3·5H2O (1000 g, 4.03 mmol) in water (10 mL) and diethyl ether (15 mL) were added. The organic layer was separated; the water layer was washed with diethyl ether (25 mL). The organic phases were combined, dried over Na2SO4, and concentrated under reduced pressure. The crude product was purified by column chromatography on silica using petroleum ether as an eluent to give a pale-grey crystalline solid of 1-I-1,2-C2B10H11 (160 mg, yield 59%).

3.5. X-ray Diffraction Study

Single crystal X-ray diffraction experiments were carried out using a SMART APEX2 CCD diffractometer (λ(Mo-Kα) = 0.71073 Å, graphite monochromator, ω-scans) at 120 K. Collected data were processed by the SAINT and SADABS programs incorporated into the APEX2 program package [70]. The structures were solved by the direct methods and refined by the full-matrix least-squares procedure against F2 in anisotropic approximation. The refinement was carried out with the SHELXTL program [71]. The CCDC numbers (2070233 for 1,12-I2-C2B10H10 and 2074102 for 1-I-C2B10H11) contain the supplementary crystallographic data for this paper. These data can be found in the Supplementary Materials or obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, accessed on 15 March 2021.
Crystallographic data for 1,12-I2-1,2-C2B10H10: C2H10B10I2 are monoclinic, space group P21/c: a = 7.1919(8)Å, b = 15.8202(17)Å, c = 11.1509(12)Å, β = 108.809(2)°, V = 1201.0(2) Å3, Z = 4, M = 396.00, dcryst = 2.190 g⋅cm−3. wR2 = 0.0365 calculated on F2hkl for all 3557 independent reflections with 2θ < 60.4°, (GOF = 1.067, R = 0.0161 calculated on Fhkl for 3314 reflections with I > 2σ(I)).
Crystallographic data for 1-I-1,2-C2B10H11: C2H11B10I are orthorhombic, space group Pnma: a = 13.8323(9)Å, b = 8.9644(6)Å, c = 8.4539(5)Å, V = 1048.27(12) Å3, Z = 4, M = 270.11, dcryst = 1.711 g·cm−3. wR2 = 0.0747 calculated on F2hkl for all 1344 independent reflections with 2θ < 56.1°, (GOF = 1.143, R = 0.0327 calculated on Fhkl for 1203 reflections with I > 2σ(I)).

3.6. Quantum Chemical Calculations

All quantum chemical calculations were carried out with the Gaussian09 program [72]. The PBE0 functional with the triple zeta basis set was found to be reliable for the calculation of noncovalent intra- and intermolecular interactions [73,74,75] and was adopted throughout this study. Initial geometries for the optimization of all dimers considered in this study were based on the X-ray structure of a dihalogen-bonded dimer of 1,12-I2-C2B10H10 (symmetry code is 1 − x, −0.5 + y, 0.5 − z). All dimeric associates were fully optimized and converged to the energy minima. Theoretical electron density was treated within the AIM approach [76] using the AIMAll program package [77]. For energy (E) estimation, we used the E = 1/2V(r) formula [78,79], in which V(r) is the potential energy density at the bond critical point between interacting atoms. It has frequently been shown that this approach to describe noncovalent interactions demonstrates realistic energetic characteristics [80,81,82].

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cryst11040396/s1, Figure S1: Asymmetric part of 1-I-1,2-C2B10H11 molecule showing numbering scheme and the disorder of the C2/B4 atoms; Figure S2: General view of 1-I-1,2-C2B10H11; Figure S3: Halogen bonded dimer of 1-I-1,2-C2B10H11, and complete crystallographic data (cif-files) for compounds 1-I-1,2-C2B10H11 and 1,12-I2-1,2-C2B10H10.

Author Contributions

Synthesis and NMR spectroscopy study, S.A.A.; quantum-chemical calculations, A.A.A.; X-ray diffraction experiment and final manuscript writing, K.Y.S.; general manuscript concept and final manuscript writing, I.B.S.; technical editing, V.I.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Russian Science Foundation—synthesis by Grant No. 19-73-00353 and crystal packing analysis and quantum-chemical calculations by Grant No. 19-13-00238.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The supporting data are available on request from the authors.

Acknowledgments

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

Conflicts of Interest

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

References

  1. Grimes, R.N. Carboranes, 3rd ed.; Academic Press: London, UK, 2016; 1042p. [Google Scholar] [CrossRef]
  2. Kaszynski, P.; Douglass, A.G. Organic derivatives of closo-boranes: A new class of liquid crystal materials. J. Organomet. Chem. 1999, 581, 28–38. [Google Scholar] [CrossRef]
  3. Pecyna, J.; Pociecha, D.; Kaszynski, P. Zwitterionic pyridinium derivatives of [closo-1- CB9H10]- and [closo-1-CB11H12]- as high Δε additives to a nematic host. J. Mater. Chem. C 2014, 2, 1585–1591. [Google Scholar] [CrossRef] [Green Version]
  4. Pecyna, J.; Kaszyński, P.; Ringstrand, B.; Pociecha, D.; Pakhomov, S.; Douglass, A.G.; Young, V.G. Synthesis and characterization of quinuclidinium derivatives of the [closo-1-CB11H12]-anion as potential polar components of liquid crystal materials. Inorg. Chem. 2016, 55, 40167–44025. [Google Scholar] [CrossRef]
  5. Pecyna, J.; Jankowiak, A.; Pociecha, D.; Kaszyński, P. o-Carborane derivatives for probing molecular polarity effects on liquid crystal phase stability and dielectric behavior. J. Mater. Chem. C 2015, 3, 11412–11422. [Google Scholar] [CrossRef]
  6. Allis, D.G.; Spencer, J.T. Polyhedral-based nonlinear optical materials. 2. Theoretical investigation of some new high non-linear optical response compounds involving polyhedral bridges with charged aromatic donors and acceptors. Inorg. Chem. 2001, 40, 3373–3380. [Google Scholar] [CrossRef]
  7. Wang, H.-Q.; Wang, L.; Li, R.-R.; Ye, J.-T.; Chen, Z.-Z.; Chen, H.; Qiu, Y.-Q.; Xie, H.-M. Second-order nonlinear optical properties of carboranylated square-planar Pt(II) zwitterionic complexes: One-/two-dimensional difference and substituent effect. J. Phys. Chem. A 2016, 120, 9330–9340. [Google Scholar] [CrossRef] [PubMed]
  8. Jiang, P.; Wang, Z.; Moxey, G.J.; Morshedi, M.; Barlow, A.; Wang, G.; Quintana, C.; Zhang, C.; Cifuentes, M.P.; Humphrey, M.G. Syntheses and quadratic nonlinear optical properties of 2,7-fluorenylene- and 1,4-phenylene-functionalized o-carboranes. Dalton Trans. 2019, 48, 12549–12559. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, H.-Q.; Ye, J.-T.; Zhang, Y.; Zhao, Y.-Y.; Qiu, Y.-Q. A thorough understanding of the nonlinear optical properties of BODIPY/carborane/diketopyrrolopyrrole hybrid chromophores: Module contribution, linear combination, one-/two-dimensional difference and carborane’s arrangement. J. Mater. Chem. C 2019, 7, 7531–7547. [Google Scholar] [CrossRef]
  10. Wedge, T.J.; Hawthorne, M. Multidentate carborane-containing Lewis acids and their chemistry: Mercuracarborands. Coord. Chem. Rev. 2003, 240, 111–128. [Google Scholar] [CrossRef] [Green Version]
  11. Scholz, M.; Hey-Hawkins, E. Carbaboranes as Pharmacophores: Properties, Synthesis, and Application Strategies. Chem. Rev. 2011, 111, 7035–7062. [Google Scholar] [CrossRef] [PubMed]
  12. Endo, Y. Carboranes as hydrophobic pharmacophores: Applications for design of nuclear receptor ligands. In Boron-Based Compounds: Potential and Emerging Applications in Medicine; Hey-Hawkins, E., Viñas Teixidor, C., Eds.; John Wiley & Sons Ltd.: Oxford, UK, 2018; pp. 3–19. [Google Scholar] [CrossRef]
  13. Yang, X.; Jiang, W.; Knobler, C.B.; Hawthorne, M.F. Rigid-rod molecules: Carborods. Synthesis of tetrameric p-carboranes and the crystal structure of bis(tri-n-butylsilyl)tetra-p-carborane. J. Am. Chem. Soc. 1992, 114, 9719–9721. [Google Scholar] [CrossRef]
  14. Schöberl, U.; Magnera, T.F.; Harrison, R.M.; Fleischer, F.; Pflug, J.L.; Schwab, P.F.H.; Meng, X.; Lipiak, D.; Noll, B.C.; Allured, V.S.; et al. Toward a Hexagonal Grid Polymer: Synthesis, Coupling, and Chemically Reversible Surface-Pinning of the Star Connectors, 1,3,5-C6H3(CB10H10CX)3. J. Am. Chem. Soc. 1997, 119, 3907–3917. [Google Scholar] [CrossRef]
  15. Fox, M.A.; Cameron, A.M.; Low, P.J.; Paterson, M.A.J.; Batsanov, A.S.; Goeta, A.E.; Rankin, D.W.H.; Robertson, H.E.; Schirlin, J.T. Synthetic and structural studies on C-ethynyl- and C-bromo-carboranes. Dalton Trans. 2006, 3544–3560. [Google Scholar] [CrossRef] [PubMed]
  16. Safronov, A.V.; Sevryugina, Y.V.; Pichaandi, K.R.; Jalisatgi, S.S.; Hawthorne, M.F. Synthesis of closo- and nido-biscarboranes with rigid unsaturated linkers as precursors to linear metallacarborane-based molecular rods. Dalton Trans. 2014, 43, 4969–4977. [Google Scholar] [CrossRef]
  17. Himmelspach, A.; Warneke, J.; Schäfer, M.; Hailmann, M.; Finze, M. Salts of the dianions [Hg(12-X-closo-1-CB11H10)2]2- (X = I, C≡CH, C≡CFc, C≡CSiiPr3): Synthesis and spectroscopic and structural characterization. Organometallics 2015, 34, 462–469. [Google Scholar] [CrossRef]
  18. Zhang, K.; Shen, Y.; Liu, J.; Spingler, B.; Duttwyler, S. Crystal structure of a carborane endo/exo-dianion and its use in the synthesis of ditopic ligands for supramolecular frameworks. Chem. Commun. 2017, 54, 1698–1701. [Google Scholar] [CrossRef] [PubMed]
  19. Farha, O.K.; Spokoyny, A.M.; Mulfort, K.L.; Hawthorne, M.F.; Mirkin, C.A.; Hupp, J.T. Synthesis and Hydrogen Sorption Properties of Carborane Based Metal−Organic Framework Materials. J. Am. Chem. Soc. 2007, 129, 12680–12681. [Google Scholar] [CrossRef]
  20. 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, 4135–4137. [Google Scholar] [CrossRef]
  21. Bae, Y.-S.; Spokoyny, A.M.; Farha, O.K.; Snurr, R.Q.; Hupp, J.T.; Mirkin, C.A. Separation of gas mixtures using Co(II) car-borane-based porous coordination polymers. Chem. Commun. 2010, 46, 3478–3480. [Google Scholar] [CrossRef]
  22. 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] [Green Version]
  23. Andrews, P.; Hardie, M.J.; Raston, C.L. Supramolecular assemblies of globular main group cage species. Coord. Chem. Rev. 1999, 189, 169–198. [Google Scholar] [CrossRef]
  24. Hardie, M.J.; Raston, C.L. Crystalline hydrogen bonded complexes of o-carborane. CrystEngComm 2001, 3, 162–164. [Google Scholar] [CrossRef]
  25. Lo, R.; Fanfrlík, J.; Lepšík, M.; Hobza, P. The properties of substituted 3D-aromatic neutral carboranes: The potential for σ-hole bonding. Phys. Chem. Chem. Phys. 2015, 17, 20814–20821. [Google Scholar] [CrossRef]
  26. Alkorta, I.; Elguero, J.; Oliva-Enrich, J.M. Hydrogen vs. Halogen Bonds in 1-Halo-Closo-Carboranes. Materials 2020, 13, 2163. [Google Scholar] [CrossRef] [PubMed]
  27. Glukhov, I.V.; Lyssenko, K.A.; Antipin, M.Y. Crystal packing of 8,9,10,12-tetrafluoro-o-carborane: H. F versus H. H contacts. Struct. Chem. 2007, 18, 465–469. [Google Scholar] [CrossRef]
  28. Fanfrlík, J.; Holub, J.; Růžičková, Z.; Řezáč, J.; Lane, P.D.; Wann, D.A.; Hnyk, D.; Růžička, A.; Hobza, P. Competition between Halogen, Hydrogen and Dihydrogen Bonding in Brominated Carboranes. ChemPhysChem 2016, 17, 3373–3376. [Google Scholar] [CrossRef]
  29. Batsanov, A.S.; Fox, M.A.; Howard, J.A.K.; Hughes, A.K.; Johnson, A.L.; Martindale, S.J. 9,12-diiodo-1,2-dicarba-closo-dodecaborane(12). Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2003, 59, o74–o76. [Google Scholar] [CrossRef] [PubMed]
  30. Barberà, G.; Viñas, C.; Teixidor, F.; Rosair, G.M.; Welch, A.J. Self-assembly of carborane molecules via C-H⋯I hydrogen bonding: The molecular and crystal structures of 3-I-1,2-closo-C2B10H11. J. Chem. Soc. Dalton Trans. 2002, 19, 3647–3648. [Google Scholar] [CrossRef]
  31. Vaca, A.; Teixidor, F.; Kivekäs, R.; Sillanpää, R.; Viñas, C. A solvent-free regioselective iodination route of ortho-carboranes. Dalton Trans. 2006, 41, 4884–4885. [Google Scholar] [CrossRef]
  32. Ramachandran, B.M.; Knobler, C.B.; Hawthorne, M.F. Synthesis and structural characterization of symmetrical clo-so-4,7-I2-1,2- C2B10H10 and [(CH3)3NH][nido-2,4-I2-7,8-C2B9H10]. Inorg. Chem. 2006, 45, 336–340. [Google Scholar] [CrossRef]
  33. Barberà, G.; Vaca, A.; Teixidor, F.; Sillanpää, R.; Kivekäs, R.; Viñas, C. Designed synthesis of new ortho-carborane derivatives: From mono- to polysubstituted frameworks. Inorg. Chem. 2008, 47, 7309–7316. [Google Scholar] [CrossRef]
  34. Puga, A.V.; Teixidor, F.; Sillanpää, R.; Kivekäs, R.; Viñas, C. Iodinated ortho-carboranes as versatile building blocks to design intermolecular interactions in crystal lattices. Chem. Eur. J. 2009, 15, 9764–9772. [Google Scholar] [CrossRef]
  35. Bregadze, V.I. Dicarba-closo-dodecaboranes C2B10H12 and their derivatives. Chem. Rev. 1992, 92, 209–223. [Google Scholar] [CrossRef]
  36. Clark, T.; Hennemann, M.; Murray, J.S.; Politzer, P. Halogen bonding: The σ-hole. J. Mol. Model. 2007, 13, 291–296. [Google Scholar] [CrossRef]
  37. Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The Halogen Bond. Chem. Rev. 2016, 116, 2478–2601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Kolář, M.H.; Hobza, P. Computer Modeling of Halogen Bonds and Other σ-Hole Interactions. Chem. Rev. 2016, 116, 5155–5187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Gilday, L.C.; Robinson, S.W.; Barendt, T.A.; Langton, M.J.; Mullaney, B.R.; Beer, P.D. Halogen Bonding in Supramolecular Chemistry. Chem. Rev. 2015, 115, 7118–7195. [Google Scholar] [CrossRef]
  40. Kolář, M.; Hostaš, J.; Hobza, P. The strength and directionality of a halogen bond are co-determined by the magnitude and size of the σ-hole. Phys. Chem. Chem. Phys. 2014, 16, 9987–9996. [Google Scholar] [CrossRef] [PubMed]
  41. Ivanov, D.M.; Kinzhalov, M.A.; Novikov, A.S.; Ananyev, I.V.; Romanova, A.A.; Boyarskiy, V.P.; Haukka, M.; Kukushkin, V.Y. H2C(X)–X···X– (X = Cl, Br) Halogen Bonding of Dihalomethanes. Cryst. Growth Des. 2017, 17, 1353–1362. [Google Scholar] [CrossRef] [Green Version]
  42. Suponitsky, K.Y.; Burakov, N.; Kanibolotsky, A.L.; Mikhailov, V.A. Multiple Noncovalent Bonding in Halogen Complexes with Oxygen Organics. I. Tertiary Amides. J. Phys. Chem. A 2016, 120, 4179–4190. [Google Scholar] [CrossRef]
  43. Zakharkin, L.I.; Zhigareva, G.G.; Kazantsev, A.V. Some reactions of barene Gringard reagents. Zh. Obshch. Khim. 1968, 38, 89–92. [Google Scholar]
  44. Zakharkin, L.I.; Podvisotskaya, L.S. Cleavage of 1,2-dihalobarenes by alcohols to C,C?-dihalodicarbaundecaboranes (13). Russ. Chem. Bull. 1966, 15, 742. [Google Scholar] [CrossRef]
  45. Tupchauskas, A.P.; Stanko, V.I.; Ustynyuk, Y.A.; Khrapov, V.V. 1H-{11B} heteronuclear double resonance spectra of ortho-, meta-, and para-carboranes and some of their organotin derivatives. J. Struct. Chem. 1973, 13, 772–776. [Google Scholar] [CrossRef]
  46. Zakharkin, L.; Kovredov, A.; Ol’Shevskaya, V.; Shaugumbekova, Z. Synthesis of B-organo-substituted 1,2-, 1,7-, and 1,12-dicarbaclosododecarboranes(12). J. Organomet. Chem. 1982, 226, 217–222. [Google Scholar] [CrossRef]
  47. Zakharkin, L.I.; Ol’Shevskaya, V.A.; Nesmeyanov’S, A.N. Synthesis of 9-Organyl-1,2 and 1,7-Dicarba-closo-dodecaboranss(12) via the Cross-Coupling Reactions between Organozinc Compounds and 9-Iodo-1,2- or 1,7-Dicarba-closo-dodecaboranes. Synth. React. Inorg. Met. Chem. 1991, 21, 1041–1046. [Google Scholar] [CrossRef]
  48. Zheng, Z.; Jiang, W.; Zinn, A.A.; Knobler, C.B.; Hawthorne, M.F. Facile Electrophilic Iodination of Icosahedral Carboranes. Synthesis of Carborane Derivatives with Boron-Carbon Bonds via the Palladium-Catalyzed Reaction of Diiodocarboranes with Grignard Reagents. Inorg. Chem. 1995, 34, 2095–2100. [Google Scholar] [CrossRef]
  49. Viñas, C.; Barbera, G.; Oliva, J.M.; Teixidor, F.; Welch, A.J.; Rosair, G.M. Are halocarboranes suitable for substitution reactions? The case for 3-I-1,2-closo-C2B10H11: Molecular orbital calculations, aryldehalogenation reactions, 11B NMR interpretation of closo-carboranes, and molecular structures of 1-Ph-3-Br-1,2-closo-C2B10H10 and 3-Ph-1,2-closo-C2B10H11. Inorg. Chem. 2001, 40, 6555–6562. [Google Scholar] [CrossRef] [PubMed]
  50. Endo, Y.; Aizawa, K.; Ohta, K. Synthesis of 3-Aryl-1,2-dicarba-closo-dodecaboranes by Suzuki-Miyaura Coupling Reaction. Heterocycles 2010, 80, 369–377. [Google Scholar] [CrossRef]
  51. Anderson, K.P.; Mills, H.A.; Mao, C.; Kirlikovali, K.O.; Axtell, J.C.; Rheingold, A.L.; Spokoyny, A.M. Improved synthesis of icosahedral carboranes containing exopolyhedral B C and C C bonds. Tetrahedron 2019, 75, 187–191. [Google Scholar] [CrossRef]
  52. Anufriev, S.A.; Shmal’ko, A.V.; Suponitsky, K.Y.; Sivaev, I.B. One-pot synthesis of B-aryl carboranes with sensitive functional groups using sequential cobalt- and palladium-catalyzed reactions. Catalysts 2020, 10, 1348. [Google Scholar] [CrossRef]
  53. Janoušek, Z.; Hilton, C.L.; Schreiber, P.J.; Michl, J. C-Halogenation of the closo-[CB11H12]- Anion. Collect. Czechoslov. Chem. Commun. 2002, 67, 1025–1034. [Google Scholar] [CrossRef]
  54. Šembera, F.; Plutnar, J.; Higelin, A.; Janoušek, Z.; Císařova, I.; Michl, J. Metal complexes with very large dipole moments: The anionic carborane nitriles 12-NC-CB11X11- (X = H, F, CH3) as ligands on Pt(II) and Pd(II). Inorg. Chem. 2016, 55, 3797–3806. [Google Scholar] [CrossRef]
  55. Estrada, J.; Lugo, C.A.; McArthur, S.G.; Lavallo, V. Inductive effects of 10 and 12-vertex closo-carborane anions: Cluster size and charge make a difference. Chem. Commun. 2016, 52, 1824–1826. [Google Scholar] [CrossRef] [PubMed]
  56. Plešek, J.; Hanslík, T. Chemistry of boranes. XXIX. The synthesis of isomeric 1,9- and 1,12-dibromo- l,2-dicarba-closo-dodecaboranes. Collect. Czechoslov. Chem. Commun. 1973, 38, 335–337. [Google Scholar] [CrossRef]
  57. Anufriev, S.A.; Sivaev, I.B.; Bregadze, V.I. Synthesis of 9,9´,12,12´-substituted cobalt bis(dicarbollide) derivatives. Russ. Chem. Bull. 2015, 64, 712–717. [Google Scholar] [CrossRef]
  58. Havránek, M.; Samsonov, M.A.; Holub, J.; Ružickova, Z.; Drož, L.; Ružicka, A.; Fanfrlík, J.; Hnyk, D. The influence of halogenated hypercarbon on crystal packing in the series of 1-Ph-2-X-1,2-dicarba-closo-dodecaboranes (X = F, Cl, Br, I). Molecules 2020, 25, 1200. [Google Scholar] [CrossRef] [Green Version]
  59. Dmitrienko, A.O.; Karnoukhova, V.A.; Potemkin, A.A.; Struchkova, M.I.; Kryazhevskikh, I.A.; Suponitsky, K.Y. The influence of halogen type on structural features of compounds containing α-halo-α,α-dinitroethyl moieties. Chem. Heterocycl. Comp. 2017, 53, 532–539. [Google Scholar] [CrossRef]
  60. Wolff, M.; Okrut, A.; Feldmann, C. [(Ph)3PBr][Br7], [(Bz)(Ph)3P]2[Br8], [(n-Bu)3MeN]2[Br20], [C4MPyr]2[Br20], and [(Ph)3PCl]2[Cl2I14]: Extending the Horizon of Polyhalides via Synthesis in Ionic Liquids. Inorg. Chem. 2011, 50, 11683–11694. [Google Scholar] [CrossRef]
  61. Sonnenberg, K.; Mann, L.; Redeker, F.A.; Schmidt, B.; Riedel, S. Polyhalogen and Polyinterhalogen Anions from Fluorine to Iodine. Angew. Chem. Int. Ed. 2020, 59, 5464–5493. [Google Scholar] [CrossRef]
  62. Allen, F.H.; Kennard, O.; Watson, D.G.; Brammer, L.; Orpen, A.G.; Taylor, R. Tables of bond lengths determined by X-ray and neutron diffraction. Part 1. Bond lengths in organic compounds. J. Chem. Soc. Perkin Trans. 1987, 2, S1–S19. [Google Scholar] [CrossRef]
  63. Varadwaj, P.R.; Varadwaj, A.; Marques, H.M. Halogen bonding: A halogen-centered noncovalent interaction yet to be understood. Inorganics 2019, 7, 40. [Google Scholar] [CrossRef] [Green Version]
  64. Suponitsky, K.Y.; Tsirelson, V.G.; Feil, D. Electron-density-based calculations of intermolecular energy: Case of urea. Acta Crystallogr. Sect. A Found. Crystallogr. 1999, 55, 821–827. [Google Scholar] [CrossRef] [PubMed]
  65. Suponitsky, K.Y.; Smol’Yakov, A.F.; Ananyev, I.V.; Khakhalev, A.V.; Gidaspov, A.A.; Sheremetev, A.B. 3,4-Dinitrofurazan: Structural Nonequivalence of ortho -Nitro Groups as a Key Feature of the Crystal Structure and Density. ChemistrySelect 2020, 5, 14543–14548. [Google Scholar] [CrossRef]
  66. Dalinger, I.L.; Suponitsky, K.Y.; Pivkina, A.N.; Sheremetev, A.B. Novel melt-castable energetic pyrazole: A pyrazol-yl-furazan framework bearing five nitro groups. Prop. Explos. Pyrotech. 2016, 41, 789–792. [Google Scholar] [CrossRef]
  67. Andrews, J.S.; Zayas, J.; Jones, M. 9-Iodo-o-carborane. Inorg. Chem. 1985, 24, 3715–3716. [Google Scholar] [CrossRef]
  68. Itatani, H.; Bailar, J.C. Homogenous catalysis in the reactions of olefinic substances. V. Hydrogenation of soybean oil methyl ester with triphenylphosphine and triphenylarsine palladium catalysts. J. Am. Oil Chem. Soc. 1967, 44, 147–151. [Google Scholar] [CrossRef]
  69. Armarego, W.L.F.; Chai, C.L.L. Purification of Laboratory Chemicals, 6th ed.; Butterworth-Heinemann: Burlington, NJ, USA, 2009. [Google Scholar] [CrossRef]
  70. Bruker AXS. APEX2 and SAINT; Bruker AXS Inc.: Madison, WI, USA, 2014. [Google Scholar]
  71. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  72. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A.; Kudin, K.N., Jr.; Burant, J.C.; Millam, J.M.; et al. Gaussian 03, Revision E.01; Gaussian, Inc.: Wallingford, UK, 2004. [Google Scholar]
  73. Anufriev, S.A.; Sivaev, I.B.; Suponitsky, K.Y.; Godovikov, I.A.; Bregadze, V.I. Synthesis of 10-methylsulfide and 10-alkylmethylsulfonium nido-carborane derivatives: B–H⋯π Interactions between the B–H–B hydrogen atom and alkyne group in 10-RC≡CCH2S(Me)-7,8-C2B9H11. Eur. J. Inorg. Chem. 2017, 38, 4436–4443. [Google Scholar] [CrossRef] [Green Version]
  74. Anufriev, S.A.; Sivaev, I.B.; Suponitsky, K.Y.; Bregadze, V.I. Practical synthesis of 9-methylthio-7,8-nido-carborane [9-MeS-7,8-C2B9H11]-. Some evidences of BH···X hydride-halogen bonds in 9- XCH2(Me)S-7,8-C2B9H11 (X = Cl, Br, I). J. Organomet. Chem. 2017, 849–850, 315–323. [Google Scholar] [CrossRef]
  75. Suponitsky, K.Y.; Masunov, A.E. Supramolecular step in design of nonlinear optical materials: Effect of π⋯π stacking aggregation on hyperpolarizability. J. Chem. Phys. 2013, 139, 094310. [Google Scholar] [CrossRef]
  76. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Clarendon Press: Oxford, UK, 1990. [Google Scholar]
  77. Keith, T.A. AIMAll; Version 15.05.18; TK Gristmill Software: Overland Park, KS, USA, 2015. [Google Scholar]
  78. Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen bond strengths revealed by topological analyses of experimentally observed electron densities. Chem. Phys. Lett. 1998, 285, 170–173. [Google Scholar] [CrossRef]
  79. Espinosa, E.; Alkorta, I.; Rozas, I.; Elguero, J.; Molins, E. About the evaluation of the local kinetic, potential and total energy densities in closed-shell interactions. Chem. Phys. Lett. 2001, 336, 457–461. [Google Scholar] [CrossRef]
  80. Suponitsky, K.Y.; Lyssenko, K.A.; Antipin, M.Y.; Aleksandrova, N.S.; Sheremetev, A.B.; Novikova, T.S. 4,4′-Bis(nitramino)azofurazan and its salts. Study of molecular and crystal structure based on X-ray and quantum chemical data. Russ. Chem. Bull. 2009, 58, 2129–2136. [Google Scholar] [CrossRef]
  81. Lyssenko, K.A. Analysis of supramolecular architectures: Beyond molecular packing diagrams. Mendeleev Commun. 2012, 22, 1–7. [Google Scholar] [CrossRef]
  82. Suponitsky, K.Y.; Lyssenko, K.A.; Ananyev, I.V.; Kozeev, A.M.; Sheremetev, A.B. Role of weak intermolecular interactions in the crystal structure of tetrakis-furazano[3,4-c:3′,4′-g:3″,4″-k:3"‘,4"‘-o][1,2,5,6,9,10,13,14]octaazacyclohexadecine and its solvates. Cryst. Growth Des. 2014, 14, 4439–4449. [Google Scholar] [CrossRef]
Figure 1. General view of 1,12-I2-1,2-C2B10H10. Thermal ellipsoids are given at 50% probability level.
Figure 1. General view of 1,12-I2-1,2-C2B10H10. Thermal ellipsoids are given at 50% probability level.
Crystals 11 00396 g001
Figure 2. Crystal packing fragment of 1,12-I2-1,2-C2B10H10. Halogen bonded chains are formed along axis b.
Figure 2. Crystal packing fragment of 1,12-I2-1,2-C2B10H10. Halogen bonded chains are formed along axis b.
Crystals 11 00396 g002
Figure 3. Noncovalent bonding in dimers of 1,12-I2-1,2-C2B10H10 (left) and 1,12-Br2-1,2-C2B10H10 (right).
Figure 3. Noncovalent bonding in dimers of 1,12-I2-1,2-C2B10H10 (left) and 1,12-Br2-1,2-C2B10H10 (right).
Crystals 11 00396 g003
Figure 4. Noncovalent bonding in dimers of 1,9-I2-1,2-C2B10H10 (left) and 1,3-I2-1,2-C2B10H10 (right).
Figure 4. Noncovalent bonding in dimers of 1,9-I2-1,2-C2B10H10 (left) and 1,3-I2-1,2-C2B10H10 (right).
Crystals 11 00396 g004
Table 1. Characteristics of intermolecular noncovalent interactions for dimers of 1,12-I2-1,2-C2B10H10 and 1,12-Br2-1,2-C2B10H10.
Table 1. Characteristics of intermolecular noncovalent interactions for dimers of 1,12-I2-1,2-C2B10H10 and 1,12-Br2-1,2-C2B10H10.
Distance in Å or Angle in Deg.Energy in kcal/mol
1,12-I2-1,2-
C2B10H10 (X-ray)
1,12-I2-1,2-
C2B10H10 (calc)
1,12-Br2-1,2-
C2B10H10 (calc)
1,12-I2-1,2-
C2B10H10 (calc)
1,12-Br2-1,2-
C2B10H10 (calc)
X12X1’3.5687(9)3.4553.704−2.9−1.0
B12-X12⋯X1’92.98(12)94.391.1
X12⋯X1’-C1’H7⋯X1’172.61(11)175.9147.6
H11⋯X1’3.58(2)3.513.25−0.5−0.7
H7⋯H5’3.58(2)3.523.37−0.5−0.5
H11⋯H4’--2.67-−0.5
X12X1’--2.61-−0.6
Table 2. Characteristics of intermolecular noncovalent interactions for halogen-bonded dimers of 1,3-I2-1,2-C2B10H10 and 1,9-I2-1,2-C2B10H10.
Table 2. Characteristics of intermolecular noncovalent interactions for halogen-bonded dimers of 1,3-I2-1,2-C2B10H10 and 1,9-I2-1,2-C2B10H10.
1,9-I2-1,2-C2B10H101,3-I2-1,2-C2B10H10
Distance or AngleEnergy Distance or AngleEnergy
I9⋯I1’3.461−2.7I3⋯I1’3.54−2.3
B9-I9⋯I1’92.7 B3-I3⋯I1’90.1
I9⋯I1’-C1’174.1 I3⋯I1’-C1’170.7
H12⋯I1’3.47−0.6H8⋯I1’3.44−0.7
H8⋯I1’3.49−0.6H10⋯I1’3.42−0.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Suponitsky, K.Y.; Anisimov, A.A.; Anufriev, S.A.; Sivaev, I.B.; Bregadze, V.I. 1,12-Diiodo-Ortho-Carborane: A Classic Textbook Example of the Dihalogen Bond. Crystals 2021, 11, 396. https://doi.org/10.3390/cryst11040396

AMA Style

Suponitsky KY, Anisimov AA, Anufriev SA, Sivaev IB, Bregadze VI. 1,12-Diiodo-Ortho-Carborane: A Classic Textbook Example of the Dihalogen Bond. Crystals. 2021; 11(4):396. https://doi.org/10.3390/cryst11040396

Chicago/Turabian Style

Suponitsky, Kyrill Yu., Alexei A. Anisimov, Sergey A. Anufriev, Igor B. Sivaev, and Vladimir I. Bregadze. 2021. "1,12-Diiodo-Ortho-Carborane: A Classic Textbook Example of the Dihalogen Bond" Crystals 11, no. 4: 396. https://doi.org/10.3390/cryst11040396

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