Next Article in Journal
Enzyme and Biological Activities of the Water Extracts from the Plants Aesculus hippocastanum, Olea europaea and Hypericum perforatum That Are Used as Folk Remedies in Turkey
Next Article in Special Issue
Blue Emitting Star-Shaped and Octasilsesquioxane-Based Polyanions Bearing Boron Clusters. Photophysical and Thermal Properties
Previous Article in Journal
Flexible and Reusable Ag Coated TiO2 Nanotube Arrays for Highly Sensitive SERS Detection of Formaldehyde
Previous Article in Special Issue
Complexes Between Adamantane Analogues B4X6 -X = {CH2, NH, O ; SiH2, PH, S} - and Dihydrogen, B4X6:nH2 (n = 1–4)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

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) †

1
APIGENEX s.r.o., Poděbradská 173/5, 190 00 Prague 9, Czech Republic
2
Department of General and Inorganic Chemistry, Faculty of Chemical Technology, University of Pardubice, Studentská 573, 532 10 Pardubice, Czech Republic
3
Institute of Inorganic Chemistry of the Czech Academy of Sciences, 250 68 Husinec-Řež, Czech Republic
4
Institute of Organic Chemistry and Biochemistry of the Czech Academy of Sciences, Flemingovo nám. 2, 166 10 Prague 6, Czech Republic
*
Authors to whom correspondence should be addressed.
Dedicated to the memory of Professor Bohumil Štíbr.
Submission received: 20 February 2020 / Revised: 3 March 2020 / Accepted: 4 March 2020 / Published: 6 March 2020

Abstract

:
Although 1-Ph-2-X-closo-1,2-C2B10H10 (X = F, Cl, Br, I) derivatives had been computed to have positive values of the heat of formation, it was possible to prepare them. The corresponding solid-state structures were computationally analyzed. Electrostatic potential computations indicated the presence of highly positive σ-holes in the case of heavy halogens. Surprisingly, the halogen•••π interaction formed by the Br atom was found to be more favorable than that of I.

Graphical Abstract

1. Introduction

Icosahedral closo-1,2-C2B10H12, known as o-carborane, was found to have the positive part of its relatively large value of the experimental dipole moment, 4.50 D [1], in the midpoint of the C-C vector. When one of the hypercarbon atoms is substituted with phenyl (Ph), the dipole moment is even increased to 4.93 D. The difference between these two values was interpreted as a mesomeric contribution to the overall dipole moment as a consequence of the electron transfer from the benzene ring towards the icosahedral cage. This 1-Ph-closo-1,2-C2B10H11 was structurally studied by the techniques of gas-phase electron diffraction and X-ray diffraction in the gas phase and solid state [2]. When the second hypercarbon of the cage is substituted with halogens, the overall dipole moments in the series of 1-Ph-2-X-closo-1,2-C2B10H10 (X = F, Cl, Br, and I, denoted here as 1, 2, 3, and 4, respectively) decrease, in the case of F and Cl by more than 1 D due to the electron-withdrawing effect of these two halogen atoms [3]. The fact that the dipole moment of Br- and I-derivatives only decreased by 0.8 and even 0.2 D, respectively, was ascribed to the close position of halogen and Ph, and also to the partially positively charged outer part of the heavy halogen atoms, known as σ-holes [4]. A σ-hole can be characterized by its magnitude, VS,max, defined as the value of the most positive electrostatic potential (ESP) of an electron density surface. The higher the VS,max value, the more favorable the forming σ-hole interactions [5]. The halogen•••π interactions have been extensively studied in the solid state and evaluated theoretically by quantum mechanical calculations [6,7,8]. Recently, they have even been observed in solution [9]. The σ-hole interactions with aryls have been paid considerable attention in boron cluster chemistry as well, e.g., the S•••π interaction appears in 1-Ph-closo-1-SB11H10 [10].
In the series of closo-SB11H11 and 12-X-closo-SB11H10 (X = Cl, Br, I), the experimentally determined dipole moments were reported to be 3.6, 5.5, 5.5, and 5.3 D [11], respectively. Note that this series differs from the 14 series in the lack of Ph substitution and in the absence of positive σ-holes on halogen atoms (relative σ-holes occur when halogen atoms are bonded to the B vertex). Within the context of halogen•••π interactions in brominated carbaboranes [12], this bonding motif has also appeared in 1-Ph-2-Br-closo-1,2-C2B10H10, whose structure has been established in another laboratory [13]. In order to obtain a deeper insight into halogen•••π interactions and crystal packing within the series of 1, 2, 3 and 4), we have prepared all of these halogenated 1-Ph-o-carboranes, crystallized them, and computationally analyzed the corresponding solid-state structures.

2. Results and Discussion

2.1. Syntheses

Compounds 1, 2, 3, and 4 were prepared by halogenating 1-phenyl-1,2-dicarba-closo-dodecaborane with N-fluorobenzenesulfonimide/benzene, Cl2/P4O10, Br2/toluene, and I2/tetrahydrofuran, respectively—as described in the literature [3]. All the compounds were identified by experimental 11B{1H} NMR spectra. The 11B NMR chemical shifts were compared with the theoretical shifts reported in ref. [3] (see Table 1). The largest difference of 3.4 ppm was found for B4 and B5 of 4. The purities of 14 were checked by analytical TLC.

2.2. Structural Characterization

The crystal structure of 3 had previously been reported by Welch as well [13]. It contained a noticeable interaction between the Br atom and the Ph ring (Br•••C contacts of 3.457 and 3.488 Å, the Br•••Phcenter separation of 3.463 Å and the C-Br•••Phcenter angle of 175.63°). We have solved the single-crystal structures of 1, 2, 3, and 4 (Figure 1 and Figure 2). The newly reported crystal structure of 3 confirms the formation of a characteristic interaction between the Br atom and the Ph ring (the Br1•••C7, Br1•••C8 and Br1•••Phcentroid separations of 3.416, 3.436, and 3.404 Å, respectively, and the C2-Br1•••Phcenter angle of 175.14°). The interaction between the I atom and the Ph ring was of comparable length (the I1B•••C7B, I1B•••C8B and I1B•••Phcenter separations of 3.553, 3.594, and 3.483 Å) but considerably more bent (C-I•••Phcenter angle of 145.00°).
The unit cell of 4 contains two independent molecules. The molecular structures of all complexes 14 are similar (Figure 1). The distances C1-C2 and C1-C3(Ph) lie within the intervals (1.620(13)–1.706(2) Å) and (1.476(16)–1.505(9) Å), respectively. The largest geometric differences are observed when the relative position of phenyl rings is analyzed. The torsion angle C2-C1-C3-C8 is 70.5(6)° and 72.0(2)° in the case of 1 and 2, respectively, 87.3(13)° for 3, and 84.9(11)° and 82.9(11)° for 4. Such differences can be explained by the features of crystal packing and the presence of intermolecular interactions C-X•••π in the case of Br- and I-derivatives, not observed in F- and Cl-derivatives (Figure 2), where usual B-H•••X and C-H•••H-B connections prevail.

2.3. Computations

2.3.1. Heat of Formation (ΔHf298)

The computed ΔHf298 values of the studied compounds and their 10-vertex analogues are summarized in Table 2. Since all the considered compounds have a high energy level (positive ΔHf298 values), their thermodynamic stability should be low as they can lose a great deal of energy by reacting to lower-energy products. The thermodynamic stability has decreased with the increasing atomic number of the halogen atom and with the reduced size of the carborane cage. Note that the positive values of the heat of formation do not necessarily mean experimental unavailability, as exemplified by e.g., closo-SB9H9 with the computed ΔHf298 value of 11.3 kcal mol−1 [14,15]. Moreover, we have computed the ΔHf298 of 26.5 kcal mol−1 for closo-1,2-C2B8H10, which was previously prepared as well [16].

2.3.2. Electrostatic Potential (ESP)

As the studied compounds are neutral and the halogen atoms are bound to a C-vertex, one can expect the heavier halogen atoms with highly positive σ-holes. The molecular ESP surfaces of the studied molecules were computed in order to validate this assumption (see Figure 3). Indeed, the top of the I and Br atoms with the VS,max values of 36.2 and 25.8 kcal mol−1, respectively, is the most positive part of the 3 and 4 molecules. In the case of 2, the σ-hole of the Cl atom has the VS,max value of 19.5 kcal mol−1. The H atoms of the Ph ring are thus more positive (VS,max value of 23.9 kcal mol−1). The F atom of 1 does not have a positive σ-hole (the VS,max value of −5.7 kcal mol−1) due to its large electronegativity. The most negative values (VS,min) of the molecular surfaces of the studied molecules are on BH(9) vertices, which are antipodal to CX(2) vertices. The VS,min values range from −14.0 to −12.9 kcal mol−1 (for 4 and 1 compounds, respectively). Besides hydridic BH vertices, the σ-holes on the heavier halogen can also interact with the π electrons of Ph rings, which have a negative ESP surface as well (the VS,min values range between −6.8 and −6.4 kcal mol−1 for 4 and 2 compounds, respectively).

2.3.3. Interactions in the Single Crystals of 14

Interactions in the reported single-crystal structures were studied by computing two-body and many-body interaction energy (ΔE2 and ΔEMB) values between the central molecules and two layers of surroundings molecules. The first layer consisted of molecules within 5 Å of the central molecule, and the second layer was formed by molecules within 5 Å of the first layer. The obtained sums of the ΔE values are summarized in Table 3. The computed total binding became more favorable with the increasing atomic number of the halogen atom (i.e. −48.4, −51.4, −53.3, and −55.2 kcal mol−1 for F-, Cl-, Br-, and I-containing compounds, respectively). The total binding thus correlated more with the molecular masses of these molecules (R2 of 0.91) than with their experimental dipole moments (R2 of 0.76).
The interaction motifs with the most favorable ΔE2 values are shown in Figure 4 and the corresponding values in Table 4. This analysis confirmed the strength of the halogen•••π interaction of 3—the motif with the C-Br•••Ph interaction had the ΔE2 of −6.91 kcal mol−1 at the DFT-D3 level. Considering that each molecule of 3 formed two such C-Br•••Ph interactions, it thus accounted for about 26% of the total computed binding of 3. According to the SAPT0 decomposition, this motif was mainly stabilized by dispersion, which formed approximately 65 of the attractive terms. The second most important term was electrostatic. It formed about 28% of the attractive terms, which was the largest contribution to the electrostatic term among all the motifs studied (see Table 4). The second most favorable motif of 3 had two diH-bonds and contributed considerably less to the overall binding. With the ΔE2 of −5.67 kcal mol−1, it only formed about 11% of the total binding of 3.
In the case of 4, the motif with the halogen•••π interaction had the ΔE2 of −5.79 kcal mol−1 and thus formed about 21% of the overall binding of 4. Therefore, it was less favorable than the motif with the halogen•••π interaction of 3. It was surprising considering the large VS,max value of 4 (see part 2.3.2.). However, examples of a reverse hierarchy in strength of halogen interactions have already been reported in literature [17,18]. In our case, the lack of strength of the I•••π interaction corresponded to the bent C-I•••Phcenter angle. An optimal arrangement for a σ-hole hole interaction is linear, whereas the C-I•••Phcenter angle was about 145° in the case of 4. Additionally, we have modeled a hypothetical dimer of 4 stabilized by a halogen•••π interaction in an optimal arrangement. The obtained motif had the I•••Phcenter separation of 3.6 Å, the C-I•••Phcenter angle of 170° and the ΔE2 of −8.50 kcal mol−1 at the DFT-D3 level, which demonstrated the capability of 4 to form a very strong iodine•••π interaction. These results indicated that crystal packing effects made the C-I•••π interaction of 4 weaker. Consequently, the most favorable crystallographic motif of 4 was the motif A•••B, which had the diH-bond with the length of 2.26 Å and the ΔE2 value of −5.98 kcal mol−1.
The most favorable motif of 2 had the highly negative ΔE2 of −7.10 kcal mol−1 (the most negative ΔE2 of this study). This motif formed about 14% of the total binding of 2 and did not have any close contact below the sum of van der Waals radii. The motif can be characterized by a large dispersion term, which formed about 73% of the attractive terms of the SAPT (Symmetry Adapted Perturbation Theory) decomposition. The second most favorable motif of 2 had multiple diH-bonds, the ΔE2 of −7.10 kcal mol−1 and a large dispersion term (i.e., about 75% of the attractive terms).
The two most favorable motifs of 1 had comparable ΔE2 of −6.93 and −6.58 kcal mol−1. Together, they formed about 28% of the total binding of 1. Neither of them formed close contact below the sum of van der Waals radii, and both had a large dispersion term in the SAPT decomposition, i.e., the dispersion term ranged from 75 to 80% of the attractive terms.

2.4. Cambridge Structural Database (CSD) search

We have searched the Cambridge structural database (CSD) [19] for X-ray structures containing halogenated carboranes that exhibit interactions between a halogen and a Ph ring. The analysis of CSD [19], however, did not show any analogous halogen•••Ph interactions in similar ortho-carborane derivatives. Additionally, we analyzed short B-X•••Ph-ring contacts in various halogenated boron compounds (for the definition of the criteria, see Figure 5). We fixed d1 and d2 to be less than sum of van der Waals radii of the appropriate elements [20], as well as angles B1-X•••C1,2 (90° < α1, α2 < 180°). A minor number of hits was excluded as clearly not suitable for the criteria of this type of interaction. The data are presented in Table 5; the found contacts may be potential candidates for studies of unusual B-X•••π interactions.
Specifically, the set of four fluorine-substituted compounds mostly contain the side-on intermolecularly interacting compounds with the B-Hal bond lying in the plane of the aromatic ring (Figure 5C). This could be attributed more or less to the non-classical C-H•••X hydrogen bond [21,22]. Nineteen relevant chlorine-substituted compounds exhibit mainly contacts contrived by C-H•••Cl or B-H•••H-C interactions, and only four structures of ionic compounds (halogenated carbadodecaborate anions are compensated by tritylium, silylium, and borinium cations), which are considered products of Cl•••π interactions [23,24,25]. A different situation occurs for brominated compounds, where the Br•••π interaction has been found to be dominant in most compounds in a set selected based on defined parameters, except for a couple of examples of type C and B (Figure 5) and boarder-line cases [26]. Surprisingly enough, only three relevant compounds have been found in the set of iodo compounds [27,28,29,30,31,32,33]. To conclude here, the most probable is an interaction of the desired type in brominated and iodinated compounds, where the aromatic ring is not a part of the same moiety as the halogen. The desired criteria are also accomplished for chlorinated ionic compounds, where the weak nucleophiles, such as CHB11X11, are compensated for by aromatic ring-containing cations. If all the restrictions were removed, leaving only the Ph ring, halogen, and three boron atoms, then 418 hits would be obtained. One can hence assume that the probability of the formation of such a motif in crystals containing both halogenated boranes and aromatic systems is 8.9%.

3. Materials and Methods

3.1. X-Ray Crystallography

The X-ray data for the compounds 14 (colorless crystals obtained by slow evaporation of a hexane solution) were collected at 150(2)K with a Bruker D8-Venture diffractometer equipped with a Mo (Mo/Kα radiation; λ = 0.71073 Å) microfocus X-ray (IµS) source, by a Photon CMOS detector and an Oxford Cryosystems cooling device. The frames were integrated with the Bruker SAINT software package [34] using a narrow-frame algorithm. The data were corrected for absorption effects using the Multi-Scan method (SADABS) [35]. The obtained data were treated by XT-version 2014/5 [36] and SHELXL-2017/1 [37] software implemented in the APEX3 v2018.1-0 (Bruker AXS Inc., Madison, WI, USA) system. Compound 1 exhibits a disorder of B(5)-H(5) and C(2)-F(1) groups (50:50). H atoms were placed in calculated positions and refined in the “riding model”. Rint = Σ|Fo2 − Fo,mean2|/ΣFo2, S = [Σ (w(Fo2 − Fc2)2)/(Ndiffrs − Nparams)]½ for all data, R(F) = Σ||Fo||Fc|||Fo| for observed data, wR(F2) = [Σ(w(Fo2 − Fc2)2)/(Σw(Fo2)2)]½ for all data. Crystallographic data for the structural analysis have been deposited with the Cambridge Crystallographic Data Centre CCDC no. 1981411–1981414. Copies of this information may be obtained free of charge from The Director, CCDC, 12 Union Road, Cambridge CB2 1EY, UK (fax: +44-1223-336033; email: [email protected] or www: http://www.ccdc.cam.ac.uk). For experimental findings see Table 6.

3.2. Computations

3.2.1. Electrostatic Potential (ESP)

The molecular ESP surfaces were computed on the 0.001 a.u. molecular surfaces at the HF/def2-TZVP level using the Gaussian09 [38] and Molekel4.3 [39,40] programs.

3.2.2. Heat of Formation (ΔHf298)

For optimization, we used the DFT/B-P86/def2QZVP level and the LBFGS algorithm with strict optimization criteria (i.e. ΔE < 0.0006 kcal mol−1, the maximal gradient <0.12 kcal mol−1 Å−1 and the RMS of gradient <0.06 kcal mol−1 Å−1. Harmonic vibrational calculations for the ZPVE and other thermodynamic contributions were computed at the DFT/B-P86/def2QZVP level. The energy calculations of the studied molecules and the atoms they constitute were performed at the DFT/B3LYP/def2QZVP level. ΔHf298 values were computed by Cuby4 [41] program package, which called Turbomole 7.0 [42] for a harmonic vibrational and energy calculations.

3.2.3. Interaction Energy

The interactions of the crystal structures were studied by using a cluster model. Hydrogen atoms of the central molecule and the surrounding molecules of the first layer were optimized by the DFT-D3/BLYP/DZVP method [43]. The resolution-of-identity (RI) approximation to the DFT method was used. Hydrogen atoms of the surrounding molecules of the second layer were optimized by the semiempirical quantum mechanical PM6-D3H4X method [44]. Heavy atoms were kept in crystallographic positions. Interaction energies were computed at the DFT-D3/TPSS/TZVPP level. Two-body interaction energy (ΔE2) was computed as the energy difference between the energy of the dimer and the sum of monomer energies. For the first layer, the interaction energy between the central molecule and the whole first layer (ΔE(AQ)) was computed as well. The many-body interaction energy (ΔEMB) was computed as the difference between ΔE(AQ) and the sum of ΔE2 values. ΔE values of selected motifs were decomposed using symmetry-adapted perturbation-theory (SAPT) methodology. The simplest truncation of SAPT (SAPT0) decomposition [45] was performed with the recommended jun-cc-pVDZ basis set [46]. Turbomole (7.0) [42], MOPAC2016 [47], PSI4 [48], and Cuby4 [41] programs were used.

4. Conclusions

A series of 14 derivatives was prepared, crystallized, and computationally analyzed. Even though their heat of formation had been computed to be positive, it was possible to prepare them. The obtained solid-state structures were computationally analyzed and the presence of σ-holes in the case of heavy halogens was computationally established. Interestingly, the halogen•••π interaction coming from the Br atom was found to be more favorable than that of I.

Author Contributions

Synthesis, M.H., L.D., and J.H.; computations, J.F. and D.H.; structural characterization and CSD search M.A.S., Z.R., and A.R. All authors have read and agreed to the published version of the manuscript.

Funding

The contribution of M.A.S. was supported from the European Social Fund—the project "International Mobility of Researchers at the University of Pardubice" (No. CZ.02.2.69/0.0/0.0/16_027/0008008)". We acknowledge the financial support from the Czech Science Foundation (project No. 19-17156S).

Acknowledgments

We thank Iegor Vyshnytskyi (APIGENEX s.r.o.) for his help with the preparation of the studied compounds.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hnyk, D.; Všetečka, V.; Drož, L.; Exner, O. Charge Distribution within 1,2-Dicarba-closo-dodecaborane: Dipole Moments of its Phenyl Derivatives. Collect. Czech. Chem. Commun. 2001, 66, 1375–1379. [Google Scholar] [CrossRef]
  2. Brain, P.T.; Cowie, J.; Donohoe, D.J.; Hnyk, D.; Rankin, D.W.H.; Reed, D.; Reid, B.D.; Robertson, H.E.; Welch, A.J. 1-Phenyl-1,2-dicarba-closo-dodecaborane, 1-Ph-1,2-closo-C2B10H11. Synthesis, Characterization, and Structure in the Gas Phase by Electron Diffraction, in the Crystalline Phase at 199 K by X-ray Diffraction, and by ab Initio Computations. Inorg. Chem. 1996, 35, 1701–1708. [Google Scholar] [PubMed]
  3. Hnyk, D.; Všetečka, V.; Drož, L. Charge Distribution within 1-Ph-2-X-1,2-dicarba-closo-dodecaboranes, (X = F, Cl, Br, I): A Dipole Moment and Computational Study. J. Mol. Struct. 2010, 978, 246–249. [Google Scholar] [CrossRef]
  4. Clark, T.; Hennemann, M.; Murray, J.S.; Politzer, P. Halogen Bonding: The σ-Hole. J. Mol. Model. 2007, 13, 291–296. [Google Scholar] [CrossRef]
  5. 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, 23279–23280. [Google Scholar] [CrossRef]
  6. Gamez, P. The anion-π interaction: Naissance and establishment of peculiar supramolecular bond. Inorg. Chem. Front. 2014, 1, 35–43. [Google Scholar] [CrossRef]
  7. Prasanna, M.D.; Guru Rown, T.N. C-halogen•••π interactions and their influence on molecular conformation and crystal packing: A database study. Cryst. Eng. 2000, 3, 135–154. [Google Scholar] [CrossRef]
  8. Youn, I.S.; Kim, D.Y.; Cho, W.J.; Madridejos, J.M.L.; Lee, H.M.; Kolaski, M.; Lee, J.; Baig, C. Halogen-π Interactions between Benzene and X2/CX4 (X = Cl, Br): Assessment of Various Density Functionals with Respect to CCSD(T). J. Phys Chem. A 2016, 120, 9305–9314. [Google Scholar] [CrossRef]
  9. Sun, H.; Horatscheck, A.; Martos, V.; Bartetzko, M.; Uhrig, U.; Lentz, D.; Schmieder, P.; Nazare, M. Direct Experimental Evidence for Halogen-Aryl π Interactions in Solution from Molecular Torsion Balances. Angew. Chem. Int. Ed. 2017, 56, 6454–6458. [Google Scholar] [CrossRef] [PubMed]
  10. Fanfrlík, J.; Přáda, A.; Padělková, Z.; Pecina, A.; Macháček, J.; Lepšík, M.; Holub, J.; Růžička, A.; Hnyk, D.; Hobza, P. The Dominant Role of Chalcogen Bonding in the Crystal Packing of 2D/3D Aromatics. Angew. Chem. Int. Ed. 2014, 53, 10139–10142. [Google Scholar] [CrossRef] [PubMed]
  11. Macháček, J.; Plešek, J.; Holub, J.; Hnyk, D.; Všetečka, V.; Císařová, I.; Kaupp, M.; Štíbr, B. New Route to 1-Thia-closo-dodecaborane(11), closo-1-SB11H11, and its Halogenation Reactions. The Effect of the Halogen on the Dipole Moments and the NMR Spectra and the Importance of Spin−Orbit Coupling for the 11B Chemical Shifts. Dalton Trans. 2006, 1024–1029. [Google Scholar]
  12. 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]
  13. McGrath, T.D.; Welch, A.J. Steric Effects in Heteroboranes. IV. 1-Ph-2-Br-1,2-closo-C2B10H10. Acta Cryst. C51 1995, 649–651. [Google Scholar] [CrossRef]
  14. Holub, J.; Vrána, J.; Růžička, A.; Růžičková, Z.; Fanfrlík, J.; Hnyk, D. Thiaboranes on Both Sides of the Icosahedral Barrier: Retaining and Breaking the Barrier with Carbon Functionalities. ChemPlusChem 2019, 84, 822–827. [Google Scholar] [CrossRef] [PubMed]
  15. Hnyk, D.; Wann, D.A.; Holub, J.; Samdal, S.; Rankin, D.W.H. Why is the Antipodal Effect in closo-1-SB9H9 so Large? A Possible Explanation Based on the Geometry from the Concerted Use of Gas Electron Diffraction and Computational Methods. Dalton Trans. 2011, 40, 5734–5737. [Google Scholar] [PubMed] [Green Version]
  16. Hnyk, D.; Rankin, D.W.H.; Robertson, H.E.; Hofmann, M.; Schleyer, P.v.R.; Bühl, M. Molecular Structure of 1,2-Dicarba-closo-decaborane(10) As Studied by the Concerted Use of Electron Diffraction and ab Initio Calculations. Inorg. Chem. 1994, 33, 4781–4786. [Google Scholar] [CrossRef]
  17. Aubert, E.; Espinossa, E.; Nicolas, I.; Jeanin, O.; Fourmigue, M. Toward a reverse hierarchy of halogen bonding between bromine and iodine. Faraday Discuss. 2017, 203, 389–406. [Google Scholar] [CrossRef] [Green Version]
  18. Ciancaleoni, G.; Belpassi, L. Disentanglement of orthogonal hydrogen and halogen bonds via natural orbital for chemical valence: A charge displacement analysis. J. Comput. Chem. 2020, 1–9. [Google Scholar] [CrossRef]
  19. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Cryst. 2016, B72, 171–179. [Google Scholar] [CrossRef]
  20. Batsanov, S.S. Van der Waals Radii of Elements. Inorganic Materials 2001, 37, 871–885. [Google Scholar] [CrossRef]
  21. Kobayashi, Y.; Popov, A.A.; Miller, S.M.; Anderson, O.P.; Strauss, S.H. Synthesis and structure of Ag(1-Me-12-SiPh3-CB11F10): Selective F12 substitution in 1-Me-CB11F11 and the first Ag(arene)4+ tetrahedron. Inorg. Chem. 2007, 46, 8505–8507. [Google Scholar] [CrossRef]
  22. Ivanova, S.M.; Ivanov, S.V.; Miller, S.M.; Anderson, O.P.; Solntsev, K.A.; Strauss, S.H. Mono-, Di-, Tri-, and Tetracarbonyls of Copper(I), Including the Structures of Cu(CO)2(1-Bn-CB11F11) and [Cu(CO)4][1-Et-CB11F11]. Inorg. Chem. 1999, 38, 3756–3757. [Google Scholar] [CrossRef]
  23. Romanato, P.; Duttwyler, S.; Linden, A.; Baldridge, K.K.; Siegel, J.S. Competition between π-arene and lone-pair halogen coordination of silylium ions? J. Am. Chem. Soc. 2011, 133, 11844–11846. [Google Scholar] [CrossRef] [PubMed]
  24. Shoji, Y.; Tanaka, N.; Mikami, K.; Uchiyama, M.; Fukushima, T. A two-coordinate boron cation featuring C-B+-C bonding. Nat. Chem. 2014, 6, 498–503. [Google Scholar] [CrossRef] [PubMed]
  25. Geis, V.; Guttsche, K.; Knapp, C.; Scherer, H.; Uzun, R. Synthesis and characterization of synthetically useful salts of the weakly-coordinating dianion [B12Cl12]2−. Dalton Trans. 2009, 2687–2694. [Google Scholar] [CrossRef] [PubMed]
  26. Kordts, N.; Kunzler, S.; Rathjen, S.; Sieling, T.; Groekappenberg, H.; Schmidtmann, M.; Muller, T. Silyl Chalconium Ions: Synthesis, Structure and Application in Hydrodefluorination Reactions. Chem. Eur. J. 2017, 23, 10068–10079. [Google Scholar] [CrossRef]
  27. Douvris, C.; Reed, C.A. Increasing the Reactivity of Vaska’s Compound. Oxidative Addition of Chlorobenzene at Ambient Temperature. Organometallics 2008, 27, 807–810. [Google Scholar]
  28. Binder, H.; Kellner, R.; Vaas, K.; Hein, M.; Baumann, F.; Wanner, M.; Winter, R.; Kaim, W.; Honle, W.; Grin, Y.; et al. The closo-cluster triad: B9X9, [B9X9](•−), and [B9X9](2−) with tricapped trigonal prisms (X = Cl, Br, I). Syntheses, crystal and electronic structures. Z. Anorg. Allg. Chem. 1999, 625, 1059–1072. [Google Scholar] [CrossRef]
  29. Gu, W.; McCulloch, B.J.; Reibenspies, J.H.; Ozerov, O.V. Improved methods for the halogenation of the [HCB11H11] anion. Chem. Commun. 2010, 46, 2820–2822. [Google Scholar] [CrossRef]
  30. Yamaguchi, T.; Fuku-en, S.; Sugawara, S.; Kojima, S.; Yamamoto, Y. Demethylation of an Allene Bearing Two Dimethoxythioxanthene Groups by Oxidation via a Vinyl Cation Intermediate. Aust. J. Chem. 2010, 63, 1638. [Google Scholar] [CrossRef]
  31. Rifat, A.; Mahon, M.F.; Weller, A.S. Dehydrogenation of cyclohexenes to cyclohexadienes by [(PPh3)2Rh]+. The isolation of an intermediate in the dehydrogenation of cyclohexane to benzene: Crystal structure of [(η4-C6H8)Rh(PPh3)2][closo-CB11H6Br6]. J. Organomet. Chem. 2003, 667, 1–4. [Google Scholar] [CrossRef]
  32. Douvris, C.; Stoyanov, E.S.; Tham, F.S.; Reed, C.A. Isolating fluorinated carbocations. Chem. Commun. 2007, 1145–1147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Moxham, G.L.; Douglas, T.M.; Brayshaw, S.K.; Kociok-Kohn, G.; Lowe, J.P.; Weller, A.S. The role of halogenated carborane monoanions in olefin hydrogenation catalysed by cationic iridium phosphine complexes. Dalton Trans. 2006, 5492–5505. [Google Scholar] [CrossRef] [PubMed]
  34. Bruker. SAINT (Version 8.38A) in APEX3 (Version 2018.1-0); Bruker AXS Inc.: Madison, WI, USA, 2018. [Google Scholar]
  35. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Cryst. 2015, 48, 3–10. [Google Scholar] [CrossRef] [Green Version]
  36. Sheldrick, G.M. SHELXT-Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  37. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  38. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision, D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  39. Flűkiger, P.; Lűthi, H.P.; Portmann, S.; Weber, J. MOLEKEL 4.3; Swiss Center for Scientific Computing: Manno, Switzerland, 2000. [Google Scholar]
  40. Portmann, S.; Luthi, H.P. MOLEKEL: An Interactive Molecular Graphic Tool. CHIMIA Int. J. Chem. 2000, 54, 766–770. [Google Scholar]
  41. Řezáč, J. Cuby: An Integrative Framework for Computational Chemistry. J. Comput. Chem. 2016, 37, 1230–1237. [Google Scholar] [CrossRef]
  42. Ahlrichs, R.; Bar, M.; Haser, M.; Horn, H.; Kolmel, C. Electronic Structure Calculations on Workstation Computers: The Program System Turbomole. Chem. Phys. Lett. 1989, 162, 165–169. [Google Scholar] [CrossRef]
  43. Hostaš, J.; Řezáč, J. Accurate DFT-D3 Calculations in a Small Basis Set. J. Chem. Theory Comput. 2017, 13, 3575–3585. [Google Scholar] [CrossRef]
  44. Řezáč, J.; Hobza, P. Advanced Corrections of Hydrogen Bonding and Dispersion for Semiempirical Quantum Mechanical Methods. J. Chem. Theory Comput. 2012, 8, 141–151. [Google Scholar] [CrossRef] [PubMed]
  45. Jeziorski, B.; Moszynski, R.; Szalewicz, K. Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of van der Waals Complexes. Chem. Rev. 1994, 94, 1887–1930. [Google Scholar] [CrossRef]
  46. Parker, T.M.; Burns, L.A.; Parrish, R.M.; Ryno, A.G.; Sherrill, C.D. Levels of Symmetry Adapted Perturbation Theory (SAPT). I. Efficiency and Performance for Interaction Energies. J. Chem. Phys. 2014, 140, 094106. [Google Scholar] [PubMed]
  47. Stewart, J.P. Optimization of parameters for semiempirical methods IV: Extension of MNDO, AM1, and PM3 to more main group elements. J. Mol. Model. 2004, 10, 155–164. [Google Scholar] [CrossRef]
  48. Turney, J.M.; Simmonett, A.C.; Parrish, R.M.; Hohenstein, E.G.; Evangelista, F.; Fermann, J.T.; Mintz, B.J.; Burns, L.A.; Wilke, J.J.; Abrams, M.L.; et al. Psi4: An Open-Source Ab Initio Electronic Structure Program. WIREs Comput. Mol. Sci. 2012, 2, 556–565. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds 14 are available from the authors.
Figure 1. The molecular structures of complexes 14. Thermal ellipsoids are at the 40% probability level.
Figure 1. The molecular structures of complexes 14. Thermal ellipsoids are at the 40% probability level.
Molecules 25 01200 g001
Figure 2. A fragment of crystal packing in 14. Thermal ellipsoids are at the 40% probability level.
Figure 2. A fragment of crystal packing in 14. Thermal ellipsoids are at the 40% probability level.
Molecules 25 01200 g002
Figure 3. The computed electrostatic potential (ESP) molecular surfaces of the 1, 2, 3, and 4 compounds. The ESP has been computed at the HF/def2TZVP level. The ESP color range is in kcal mol−1
Figure 3. The computed electrostatic potential (ESP) molecular surfaces of the 1, 2, 3, and 4 compounds. The ESP has been computed at the HF/def2TZVP level. The ESP color range is in kcal mol−1
Molecules 25 01200 g003
Figure 4. The most stable interaction motifs from the studied crystals of the 14 compounds. The positions of H atoms have been optimized at the DFT-D3/BLYP/DZVP level.
Figure 4. The most stable interaction motifs from the studied crystals of the 14 compounds. The positions of H atoms have been optimized at the DFT-D3/BLYP/DZVP level.
Molecules 25 01200 g004
Figure 5. A general scheme of the fragment used in the analysis of the CCDC and some examples of the structural motifs found. The A, B, C are the fragment used.
Figure 5. A general scheme of the fragment used in the analysis of the CCDC and some examples of the structural motifs found. The A, B, C are the fragment used.
Molecules 25 01200 g005
Table 1. 11B chemical shifts (in ppm) for 14 with respect to BF3.OEt2. 11B NMR spectra were recorded on a Varian Unity—500 instrument in CDCl3 solution. Computed shifts taken from ref. [3] are shown in parentheses. The calculations were performed at the GIAO-B3LYP/II//MP2/6-31G* level (DZP + ECP were used for 34).
Table 1. 11B chemical shifts (in ppm) for 14 with respect to BF3.OEt2. 11B NMR spectra were recorded on a Varian Unity—500 instrument in CDCl3 solution. Computed shifts taken from ref. [3] are shown in parentheses. The calculations were performed at the GIAO-B3LYP/II//MP2/6-31G* level (DZP + ECP were used for 34).
CompoundB9B12B4, B5B7, B11B3, B6B8, B10
1−6.6
(−7.1)
−11.2
(−11.3)
−12.6
(−14.0)
−13.6
(−15.6)
−14.6
(−16.1)
−14.6
(−16.4)
2−4.6
(−5.1)
−6.4
(−6.8)
−10.2
(−11.6)
−10.2
(−11.6)
−10.7
(−12.3)
−11.8
(−13.5)
3−4.2
(−4.8)
−5.3
(−6.0)
−9.2
(−11.2)
−10.8
(−12.0)
−10.8
(−12.8)
−10.8
(−13.1)
4−3.2
(−4.4)
−3.7
(−4.5)
−7.8
(−11.2)
−9.4
(−11.6)
−9.4
(−12.6)
−10.2
(−12.8)
Table 2. The computed heats of formation (ΔHf298) in kcal mol−1.
Table 2. The computed heats of formation (ΔHf298) in kcal mol−1.
CompoundΔHf298
12-vertex series
1-Ph-2-F-closo-1,2-C2B10H10 (1)2.7
1-Ph-2-Cl-closo-1,2-C2B10H10 (2)45.7
1-Ph-2-Br-closo-1,2-C2B10H10 (3)57.0
1-Ph-2-I-closo-1,2-C2B10H10 (4)64.9
10-vertex series
1-Ph-2-F-closo-1,2-C2B8H830.9
1-Ph-2-Cl-closo-1,2-C2B8H873.2
1-Ph-2-Br-closo-1,2-C2B8H884.6
1-Ph-2-I-closo-1,2-C2B8H892.6
Table 3. Two-body and many-body interaction energies (ΔE2 and ΔEMB) computed at the DFT-D3/TPSS/TZVPP level in kcal mol−1.
Table 3. Two-body and many-body interaction energies (ΔE2 and ΔEMB) computed at the DFT-D3/TPSS/TZVPP level in kcal mol−1.
CompoundΣΔE2
(1st Layer)
ΔEMB
(1st Layer)
ΣΔE2
(2nd Layer)
Total
1-Ph-2-F-closo-1,2-C2B10H10 (1)−50.745.04−2.69−48.39
1-Ph-2-Cl-closo-1,2-C2B10H10 (2)−52.295.51−4.60−51.38
1-Ph-2-Br-closo-1,2-C2B10H10 (3)−53.724.05−3.58−53.25
1-Ph-2-I-closo-1,2-C2B10H10 (4)−56.313.45−2.37−55.23
Table 4. Interaction energies computed at the DFT-D3/TPSS/TZVPP level. The interaction energies have been decomposed into electrostatic (Eelec), induction (Eind), dispersion (Edisp), and exchange (Eexch) contributions using the SAPT0/jun-cc-pVDZ methodology. All energies are in kcal mol−1. The relative values in parentheses show the contribution to the sum of all the attractive energy terms of SAPT.
Table 4. Interaction energies computed at the DFT-D3/TPSS/TZVPP level. The interaction energies have been decomposed into electrostatic (Eelec), induction (Eind), dispersion (Edisp), and exchange (Eexch) contributions using the SAPT0/jun-cc-pVDZ methodology. All energies are in kcal mol−1. The relative values in parentheses show the contribution to the sum of all the attractive energy terms of SAPT.
MotifDFT-D3SAPT0
TotalEelecEindEdispEexch
1-Ph-2-F-closo-1,2-C2B10H10 (1)
A•••B −6.93−7.20−1.92
(15.8%)
−0.57
(4.7%)
−9.65
(79.5%)
4.94
A•••C−6.58−7.15−2.56
(19.5%)
−0.68
(5.2%)
−9.87
(75.3%)
5.96
1-Ph-2-Cl-closo-1,2-C2B10H10 (2)
A•••B−7.10−7.82−3.44
(21.5%)
−0.91
(5.7%)
−11.63
(72.8%)
8.15
A•••C−5.43−4.64−1.68
(15.9%)
−1.03
(9.8 %)
−7.87
(74.4%)
5.95
1-Ph-2-Br-closo-1,2-C2B10H10 (3)
A•••B1−6.91−7.10−4.80
(27.6%)
−1.27
(7.3%)
−11.30
(65.1%)
10.23
A•••C2−5.67−5.27−2.19
(22.6%)
−0.73
(7.5%)
−6.75
(69.8%)
4.40
1-Ph-2-I-closo-1,2-C2B10H10 (4)
A•••B−5.98
A•••C−5.79
1 The analogous motif of the crystal structure by A. Welch et al. [13] was computed to have the interaction energy of −6.73 kcal mol−1 at the MP2/CBS level [12]. 2 The analogous motif of the crystal structure by Welch [13] was computed to have the interaction energy of −5.09 kcal mol−1 at the MP2/CBS level [12].
Table 5. The data obtained in the analysis of the fragment in the CCDC.
Table 5. The data obtained in the analysis of the fragment in the CCDC.
XRestraints, ÅNumber of StructuresAngles (B-Hal-C) α1, α2d1, d2, Åc, Å
F2.8 < (d1, d2) < 3.1499.08–168.802.871–3.0903.235–4.013
Cl3.1 < (d1, d2) < 3.51996.77–169.763.156–3.4973.172–4.287
Br3.1 < (d1, d2) < 3.610104.44–171.013.329–3.5863.567–4.476
I3.1 < (d1, d2) < 3.84140.21–176.253.437–3.7953.416–4.078
Table 6. The refinement information and crystallographic data for 14.
Table 6. The refinement information and crystallographic data for 14.
Compound1234
Chemical formulaC8H15B10FC8H15B10ClC8H15B10BrC8H15B10I
Formula weight238.30254.75299.21346.20
Temperature/K150(2)150(2)150(2)150(2)
Crystal systemMonoclinicMonoclinicOrthorhombicMonoclinic
Space groupP21/mP21/nPbcaP21
a/Å8.6564(9)7.2920(4)10.257(3)12.0950(8)
b/Å7.5229(7)23.9912(14)11.448(3)7.2033(5)
c/Å10.5576(11)7.7979(5)24.132(7)16.9983(14)
α/°90909090
β/°106.168(3)93.397(2)9090.453(3)
γ/°90909090
Volume/Å660.33(12)1361.80(14)2833.8(14)1480.91(19)
Z2484
ρcalc g/cm31.1991.2431.4031.553
μ/mm−10.0660.2482.8702.133
F(000)2445201184664
Crystal size/mm30.989 × 0.504 × 0.3860.414 × 0.225 × 0.1480.342 × 0.192 × 0.1500.753 × 0.416 × 0.343
Radiation typeMoKα
(λ = 0.71073 Å)
MoKα
(λ = 0.71073 Å)
MoKα
(λ = 0.71073 Å)
MoKα
(λ = 0.71073 Å)
2θ range for data collection/°2.450 to 27.9962.751 to 26.4152.606 to 24.9992.396 to 28.276
Index ranges−11 < = h < =11,
−9 < = k < = 9,
−13 < = l < = 13
−9 < = h < = 9,
−29 < = k < = 30,
−9 < = l < = 9
−12 < = h < = 12,
−13 < = k < = 12,
−28 < = l < = 28
−14 < = h < = 16,
−9 < = k < = 9,
−22 < = l < = 22
Reflections collected14986331081475821597
Independent reflections1704 [R(int) = 0.0745]2794 [R(int) = 0.0694]2439 [R(int) = 0.1169]6917 [R(int) = 0.0499]
Data/restraints/parameters1704/12/1152794/0/1722439/264/1726917/1/344
Goodness-of-fit on F21.0491.0671.1761.053
Final R indexes [I > 2σ(I)]R1 = 0.0620, wR2 = 0.1615R1 = 0.0479, wR2 = 0.1096R1 = 0.1237, wR2 = 0.2686R1 = 0.0370, wR2 = 0.0718
Largest diff. peak/hole/e Å-30.673 and −0.3930.282 and −0.3071.683 and −1.5211.635 and −1.431

Share and Cite

MDPI and ACS Style

Havránek, M.; Samsonov, M.A.; Holub, J.; Růžičková, Z.; Drož, L.; Růžička, 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. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25051200

AMA Style

Havránek M, Samsonov MA, Holub J, Růžičková Z, Drož L, Růžička 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(5):1200. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25051200

Chicago/Turabian Style

Havránek, Miroslav, Maksim A. Samsonov, Josef Holub, Zdeňka Růžičková, Ladislav Drož, Aleš Růžička, Jindřich Fanfrlík, and Drahomír Hnyk. 2020. "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 25, no. 5: 1200. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25051200

Article Metrics

Back to TopTop