Skip to content
BY 4.0 license Open Access Published by De Gruyter January 31, 2020

The crystal structure of the first ether solvate of hexaphenyldistannane [(Ph3Sn)2 • 2 THF]

  • Jonathan O. Bauer EMAIL logo

Abstract

Structural investigations of molecular crystal solvates can provide important information for the targeted crystallization of particular inclusion compounds. Here, the crystal structure of the first ether solvate of hexaphenyldistannane [(Ph3Sn)2 • 2 THF] is reported. Structural features in terms of host-guest interactions and in the context of the previously reported polymorphs and solvates of (Ph3Sn)2 are discussed.

Molecules in the solid state can arrange in more than just one crystal packing (polymorphs) and can crystallize by incorporating stoichiometric or non-stoichiometric amounts of solvent molecules in the crystal lattice to form inclusion compounds (solvates) (Bernstein, 2002; Lee et al., 2011). Usually, even the inclusion of small solvent molecules into a molecular crystal network leads to a significant change of the crystal structure (Atwood et al., 2017). Polymorphism and solvatomorphism are structural phenomenons in solid-state chemistry that can substantially affect the physicochemical properties of a crystalline molecular compound (Brittain, 2012a; Nassimbeni, 2003) and provide valuable information for crystal engineering (Desiraju, 1995). Therefore, nonsolvated polymorphs and solvates of molecular crystals have gained great importance in pharmaceutical science (Brittain, 2012b; Shan and Zaworotko, 2008). Different crystal structures and crystalline solvates of the same compound can for example influence the stability, solubility, and dissolution rate of an active pharmaceutical ingredient so that clinically relevant effects on bioavailability may result (Rexrode et al., 2019). In general, crystalline solvates can provide valuable information on the correlation between crystal structure, molecular structure, and molecule-solvent interaction. The noncovalent intermolecular interactions that determine the structure of molecular crystalline systems can vary widely in type and strength (Steed and Atwood, 2009) and range from strong ionic interactions (100-350 kJ mol-1) via hydrogen bonds (20-40 kJ mol–1), C–H···O and O–H··π interactions (2-20 kJ mol–1) through to weak nondirectional interactions like C···H in aromatic compounds (2-10 kJ mol–1) (Desiraju, 1995), which are often summarized as van der Waals forces (< 5 kJ mol–1) (Nassimbeni, 2003). Although hydrogen bonding ranging from weak to strong interactions is considered to be the most important structure-forming principle in supramolecular crystal construction (Desiraju, 2002; Steiner, 2002), there are also intriguing examples of highly stable supramolecular systems that are held together only by weak dispersive interactions and used for the interstitial inclusion of highly volatile gases like methane (Atwood et al., 2002).

In the course of our studies on the structural behaviour of triphenyl-substituted group 14 elements (Bauer and Strohmann, 2010; Bauer and Strohmann, 2015; Bauer and Strohmann, 2017; Bauer et al., 2010), we already reported on the inclusion of pyrrolidine into triphenylsilanol-based hydrogen-bonded arrangements (Bauer and Strohmann, 2015). Recently, we reported the methanol solvate of enantiomerically pure (S)-2-(triphenylsilyl)pyrrolidinium chloride (Bauer and Strohmann, 2017) as a further crystal solvate in addition to the previously described chloroform-incorporating hydrochloride (Bauer et al., 2010). Hexaphenylethane, the lightest representative within the group 14, is a hypothetical compound (Lewars, 2008) that has been the subject of intense research since its presumed discovery in 1900 (Gomberg, 1900a; Gomberg, 1900b), leading to a new conceptual understanding of the interplay between steric repulsion and noncovalent dispersive attraction (Rösel et al., 2017). Although many structural data have been collected during the last decades, single-crystals of pure hexaphenyldisilane suitable for single-crystal X-ray diffraction have not hitherto been obtained (Bernert et al., 2014; Brendler et al., 2012).

Hexaphenyldistannane, the tin homologue of hexaphenylethane, was first synthesized by Krause and Becker in 1920 by reduction of triphenyltinchloride with sodium and crystallized as the benzene solvate (Ph3Sn)2 • 2 C6H6 (Krause and Becker, 1920). The physicochemical properties of this compound have been described very carefully by the same authors and it has been found that (Ph3Sn)2 can be crystallized from m-xylene, diethyl ether, and chloroform solvent-free. The crystal structures of a monoclinic (Preut et al., 1973) and a triclinic polymorph (Dakternieks et al., 2001) of solvent-free hexaphenyldistannane have been reported. In addition to the parent phenyl-substituted compound, crystal structures of the symmetric hexaaryldistannanes (p-Tol3Sn)2 (Schneider-Koglin et al., 1991), (o-Tol3Sn)2 (Schneider-Koglin et al., 1993), and [(C6F5)3Sn]2 (Bojan et al., 2010), and two co-crystalline compounds with included hexakis(2,6-diethylphenyl)distannane (Sita and Bickerstaff, 1989) and included {[o-(CH2NMe2)C6H4] Ph2Sn}2 (Turek et al., 2009), respectively, have been reported. The dianion Sn2Ph42has been the subject of investigation in the context of phenyl-substituted Zintl anions of the heavier main-group elements (Scotti et al., 1997; Wiesler et al., 2006). To my knowledge, only for two solvates of (Ph3Sn)2 singe-crystal X-ray structure determinations have so far been performed, i.e. the trigonal compounds (Ph3Sn)2 • 2 C6H6 (Piana et al., 1991) and (Ph3Sn)2 • 2 C6H5CH3 (Eckardt et al., 1995; Karlov et al, 2013). Remarkably, thermogravimetry and X-ray powder diffraction of the benzene, toluene, fluorobenzene, chlorobenzene, and aniline solvates proved that all known crystal solvates are isomorphous and have the space group regardless of whether they are formed by symmetric or asymmetric aromatic solvents (Eckardt et al., 1995). The reason is obvious, since all asymmetric solvents in the crystal lattice are rotationally disordered leading to a highly symmetric crystallographic pseudo threefold axis (Eckardt et al., 1995). This was also observed for the toluene solvate of hexaphenyldisiloxane (Hönle et al., 1990). In addition, (Ph3Sn)2 was structurally characterized as part of a composite crystal (Lu and Cui, 2007). Hexaphenyldistannane can be considered as an interesting and simplified molecular model for investigating polymorphism and solvatomorphism in symmetrical and sterically shielded all-aryl-substituted group 14 compounds that cause crystal lattices with cavities of suitable size for the incorporation of solvent molecules.

This contribution presents the single-crystal X-ray diffraction analysis of the first crystalline ether solvate of hexaphenyldistannane [(Ph3Sn)2 • 2 THF] that was obtained by reductive Sn–Sn coupling in the presence of tert-butyllithium, presumably in an analogous manner as previously described using Grignard reagents (Dakternieks et al., 2001). (Ph3Sn)2 • 2 THF crystallized from a solvent mixture of tetrahydrofuran/diethyl ether/ pentane (2:1:1) at –30°C in the triclinic crystal system, space group as colourless needles suitable for single-crystal X-ray diffraction analysis (see Table 1 and Figure 1). The crystallographic asymmetric unit contains a half molecule of hexaphenyldistannane and one tetrahydrofuran molecule arranged around an inversion center. The two THF molecules are located at both ends of the extended Sn–Sn bond axis. The phenyl groups along the Sn–Sn axis are arranged in a gauche conformation (C–Sn–Sn–C angles: 180,00°; 58,20°; 59,96°; 61,84°), which, after comparison with the other known structures of hexaphenyldistannane, is unaffected by the packing. However, unlike the previously known solvates (space group consisting of symmetric (benzene) and asymmetric aromatic solvents (Eckardt et al., 1995; Piana et al., 1991), here the threefold axis of hexaphenyldistannane does not run perpendicularly through the THF molecules. The absence of this high symmetry in (Ph3Sn)2 • 2 THF, which in the previous cases was caused by rotational disorder, may indicate that the THF molecules are more tightly bound in the crystal lattice than the aromatic solvents. The Sn–Sn* distance with 2.7592(4) Å is quite similar to the Sn–Sn distance in the triclinic solvent-free polymorph of hexaphenyldistannane [2.7635(14) Å] (Dakternieks et al., 2001) and slightly shorter than the average Sn–Sn distance of the two independent molecules in the monoclinic form [2.759(4) Å and 2.780(4) Å] (Preut et al., 1973). The respective bond distances in the benzene and toluene solvates are 2.7666(4) Å and 2.7909(8) Å, respectively (Eckardt et al., 1995; Piana et al., 1991). The C–Sn distances are 2.138(3) Å, 2.141(3) Å, and 2.145(3) Å and are shorter than in the monoclinic solvent-free polymorph (average C–Sn distance: 2.18 Å) (Preut et al., 1973), but they are comparable to those of the other crystalline hexaphenyldistannanes and to related triphenylstannyl compounds (Bettens et al., 2009; Dakternieks et al., 1998; Reeske et al., 2001). The ring angles around the aromatic carbon atoms bound to the tin atom (approx. 118°) deviate from the ideal angle (120°) by 2°, which is known for Ph3Sn moieties (Eckardt et al., 1995). However, it has been found that this deviation is even more pronounced in triphenylsilyl-substituted derivatives (Bauer and Strohmann, 2010). The angels within the included tetrahydrofuran molecule range from almost ideal tetrahedral angles of 109.0(2) Å [C(22)–O–C(19)] to strongly distorted angles of 101.1(3) Å [C(22)–C(21)–C(20)]. Figure 2 shows some short intermolecular distances that are in the range of the sum of the van der Waals radii of the atoms involved (Rowland and Taylor, 1996). The shortest intermolecular distances can be found between the oxygen atom of the tetrahydrofuran molecule and hydrogen atoms of the phenyl rings of hexaphenyldistannane ranging from 2.628 Å [H(10)***···O] to 2.715 Å [H(16)*···O] (see Figure 2). Such weak C–H···O interactions are considered attractive and directional, and it is assumed that they can contribute significantly to the crystal packing, especially in the absence of strong hydrogen bonds (Desiraju, 1991; Desiraju, 1995). Deviations of this type of C–H···O hydrogen bonding from the directional preference of the hydrogen atoms towards the plane of the lone electron pairs have already been rationalized in the case of ethereal oxygen atoms (Desiraju, 1991; Legon and Millen, 1987). Even a short C···H distance of 2.861 Å between aryl rings can be located [H(2)—C(11)***] (see Figure 2).

Table 1

Crystal data and structure refinement of (Ph3 Sn)2 • 2 THF.

Empirical formulaC44H46O2Sn2
Formula weight [g·mol–1]844.19
Crystal systemTriclinic
Space groupP1¯
a[Å]8.5992(3)
b[Å]11.2314(3)
c[Å]11.2870(3)
α[°]61.204(3)
β[°]84.396(2)
γ[°]84.581(2)
Volume [Å3]949.34(5)
Z1
Density (calculated) ρ [g·cm–3]1.477
Absorption coefficient μ [mm–1]1.350
F(000)426
Crystal size [mm3]0.40×0.30×0.30
Theta range for data collection θ [°]2.063–25.999
Index ranges–10 ≤ h ≤ 10
–13 ≤ k ≤ 13
–13 ≤ / ≤ l3
Reflections collected24331
Independent reflections3733 (Rint= 0.0382)
Completeness to θ = 27.00°100.0%
Max. and min. transmission0.6875 and 0.6142
Data/restraints/parameters3733/0/277
Goodness-of-fit on F21.000
Final R indices [l>2σ(l)]R1 =0.0219, wR2 = 0.0606
R indices (all data)R1 =0.0227, wR2 = 0.0610
Largest diff. peak and hole [Å3]1.012 and –0.359
Figure 1 Molecular structure of (Ph3Sn)2 • 2 THF with the labeling scheme indicating the asymmetric unit (displacement ellipsoids set at the 50% probability level). Selected bond lengths [Å] and angles [°]: Sn–Sn* 2.7592(4), C(1)–Sn 2.145(3), C(7)–Sn 2.138(3), C(13)–Sn 2.141(3), C(19)–O 1.423(4), C(19)–C(20) 1.494(5), C(20)–C(21) 1.513(5), C(21)–C(22) 1.508(5), C(22)–O 1.418(4), C(6)–C(1)–C(2) 118.0(3), C(8)–C(7)–C(12) 118.1(3), C(14)–C(13)–C(18) 118.0(3), C(1)–Sn–Sn* 113.28(7), C(7)–Sn–Sn* 111.00(7), C(13)–Sn–Sn* 111.97(7), C(7)–Sn–C(1) 106.71(10), C(7)–Sn–C(13) 108.79(10), C(13)–Sn–C(1) 104.73(10), O–C(19)–C(20) 107.3(3), C(19)–C(20)–C(21) 103.5(3), C(22)–C(21)–C(20) 101.1(3), O–C(22)–C(21) 106.6(3), C(22)–O–C(19) 109.0(2). Symmetry transformations used to generate equivalent atoms: (*) –x + 1, –y + 2, –z
Figure 1

Molecular structure of (Ph3Sn)2 • 2 THF with the labeling scheme indicating the asymmetric unit (displacement ellipsoids set at the 50% probability level). Selected bond lengths [Å] and angles [°]: Sn–Sn* 2.7592(4), C(1)–Sn 2.145(3), C(7)–Sn 2.138(3), C(13)–Sn 2.141(3), C(19)–O 1.423(4), C(19)–C(20) 1.494(5), C(20)–C(21) 1.513(5), C(21)–C(22) 1.508(5), C(22)–O 1.418(4), C(6)–C(1)–C(2) 118.0(3), C(8)–C(7)–C(12) 118.1(3), C(14)–C(13)–C(18) 118.0(3), C(1)–Sn–Sn* 113.28(7), C(7)–Sn–Sn* 111.00(7), C(13)–Sn–Sn* 111.97(7), C(7)–Sn–C(1) 106.71(10), C(7)–Sn–C(13) 108.79(10), C(13)–Sn–C(1) 104.73(10), O–C(19)–C(20) 107.3(3), C(19)–C(20)–C(21) 103.5(3), C(22)–C(21)–C(20) 101.1(3), O–C(22)–C(21) 106.6(3), C(22)–O–C(19) 109.0(2). Symmetry transformations used to generate equivalent atoms: (*) –x + 1, –y + 2, –z

Figure 2 Part of the crystal structure of (Ph3Sn)2 • 2 THF illustrating short intermolecular distances found within the host-guest compound (displacement ellipsoids set at the 50% probability level). Selected bond lengths [Å] and angles [°]: H(2)···C(11)*** 2.861, C(2)–H(2)···C(ll)*** 137.85, H(4)**···0 2.706, C(4)**–H(4)**···O 146.71, H(10)***···O 2.628, C(10)***–H(10)***···O 137.02, H(16)*···O 2.715, C(16)*–H(16)*···O 153.86. Symmetry transformations used to generate equivalent atoms: (*) –x, –y +1, –z; (**) –x, –y + 1, –z +1; (***) x – 1 , y, z
Figure 2

Part of the crystal structure of (Ph3Sn)2 • 2 THF illustrating short intermolecular distances found within the host-guest compound (displacement ellipsoids set at the 50% probability level). Selected bond lengths [Å] and angles [°]: H(2)···C(11)*** 2.861, C(2)–H(2)···C(ll)*** 137.85, H(4)**···0 2.706, C(4)**–H(4)**···O 146.71, H(10)***···O 2.628, C(10)***–H(10)***···O 137.02, H(16)*···O 2.715, C(16)*–H(16)*···O 153.86. Symmetry transformations used to generate equivalent atoms: (*) –x, –y +1, –z; (**) –x, –y + 1, –z +1; (***) x – 1 , y, z

In summary, the ether solvate [(Ph3Sn)2 • 2 THF] shows that also non-aromatic guest molecules can be incorporated in the crystal lattice of hexaphenyldistannane. Remarkable are the only small deviations of the intramolecular bonding parameters in the hexaphenyldistannane molecule of the here described tetrahydrofuran solvate when compared to the other known solvatomorphs and the crystal structures of the solvent-free host compound. Evidently, the guest molecules can be exchanged over a wide structural range without requiring large changes in the packing of the host compound. Weak intermolecular C–H···O interactions may cause an additional anchoring of the entrapped ether molecules within the host structure. This may give rise to further structural model studies regarding the interstitial crystallization behaviour of all-C-substituted main group element compounds.

Experimental

General

The title compound has been obtained as a by-product during the preparation of substituted triphenylstannyl compounds from tert-butyllithium and triphenyltinchloride in a solvent mixture of tetrahydrofuran/diethyl ether/pentane (2:1:1). Colourless needles suitable for single-crystal X-ray diffraction analysis were obtained at –30°C after one day.

X-ray crystallography

Single-crystal X-ray diffraction analysis of (Ph3Sn)2 • 2 THF was performed on an Oxford Diffraction CCD Xcalibur S Diffractometer (Oxford Diffraction Ltd., 2008) at 173(2) K using graphite-monochromated Mo-Kα radiation (A = 0.71073 Å). The crystal structure was solved with direct methods [SHELXS-97 (Sheldrick, 1997; Sheldrick, 2008)] and refined against F2 with the full-matrix least-squares method [SHELXL-2018/3 (Sheldrick, 2008; Sheldrick, 2015; Sheldrick, 2018)] using the SHELX program package as implemented in WinGX (Farrugia, 2012). A multi-scan absorption correction using the implemented CrysAlis RED program was employed (Oxford Diffraction Ltd., 2008). The non-hydrogen atoms were refined using anisotropic displacement parameters. The aromatic hydrogen atoms were located on the difference Fourier map and refined independently. All other hydrogen atoms were placed in idealized geometric positions and each was assigned a fixed isotropic displacement parameter based on a riding-model with Uiso(H) = 1.2Ueq (C) for the methylene hydrogen atoms. The C–H distances constrained to 0.99 Å for the methylene groups. Crystal and refinement data are collected in Table 1. Figures 1 and 2 were created using Mercury 4.1.0 (Macrae et al., 2006). Crystallographic data of (Ph3Sn)2 • 2 THF have been deposited with The Cambridge Crystallographic Data Centre. CCDC 1955228 contains the supplementary crystallographic data for (Ph3Sn)2 • 2 THF. Copies of these data may be obtained free of charge from The Cambridge Crystallographic Data Centre (CCDC), 12 Union Road, Cambridge CB2 1EZ, UK (Fax: +44-1223-336033; e-mail: deposit@ccdc.cam.ac.uk), or via www.ccdc.cam.ac.uk/data_request/cif.

Acknowledgements

The Elite Network of Bavaria (ENB), the Bavarian State Ministry of Science and the Arts (StMWK), and the University of Regensburg are gratefully acknowledged for financial support. In addition, I thank Prof. Dr. Manfred Scheer for generous support concerning research facilities at the Institute of Inorganic Chemistry.

References

Atwood J.L., Barbour L.J., Jerga A., Storage of methane and freon by interstitial van der Waals confinement. Science, 2002, 296, 2367-2369.10.1126/science.1072252Search in Google Scholar PubMed

Atwood J.L., Gokel G.W., Barbour L.J. (Eds.), Comprehensive supramolecular chemistry II (2nd ed.). Elsevier Ltd., Amsterdam, Oxford, Cambridge, 2017, Vols. 1-9.Search in Google Scholar

Bauer J.O., Stiller J., Marqués-López E., Strohfeldt K., Christmann M., Strohmann C., Silyl-modified analogues of 2-tritylpyrrolidine: synthesis and applications in asymmetric organocatalysis. Chem. Eur. J., 2010, 16, 12553-12558.10.1002/chem.201002166Search in Google Scholar PubMed

Bauer J.O., Strohmann C., tert-Butoxytriphenylsilane. Acta Crystallogr., 2010, E66, o461-o462.10.1107/S1600536810002722Search in Google Scholar PubMed PubMed Central

Bauer J.O., Strohmann C., Hydrogen bonding principles in inclusion compounds of triphenylsilanol and pyrrolidine: Synthesis and structural features of [(Ph3SiOH)4·HN(CH24 and [Ph3SiOH·HN(CH24·CH3CO2H]. J. Organomet. Chem., 2015, 797, 52-56.10.1016/j.jorganchem.2015.07.026Search in Google Scholar

Bauer J.O., Strohmann C., Molecular structures of enantiomericallypure (S)-2-(triphenylsilyl)- and (S)-2-(methyldiphenylsilyl) pyrrolidinium salts. Inorganics, 2017, 5, 88.10.3390/inorganics5040088Search in Google Scholar

Bernert T., Winkler B., Krysiak Y., Fink L., Berger M., Alig E., et al., Determination of the crystal structure of hexaphenyldisilane from powder diffraction data and its thermodynamic properties. Cryst. Growth Des., 2014, 14, 2937-2944.10.1021/cg5002286Search in Google Scholar

Bernstein J., Polymorphism in molecular crystals. Oxford Univ. Press, New York, 2002.Search in Google Scholar

Bettens R.P.A., Dakternieks D., Duthie A., Kuan F.S., Tiekink E.R.T., The influence of crystal packing effects upon the molecular structures of Ph3Sn(CH2nSnPh3 n = 1 to 8, as determined by X-ray crystallography and DFT molecular orbital calculations. Supramolecular aggregation patterns sustained by C–H···π interactions. CrystEngComm, 2009,11, 1362-1372.10.1039/b822682bSearch in Google Scholar

Bojan R.V., López-de-Luzuriaga J.M., Monge M., Olmos M.E., Synthesis and characterization of perhalophenyltin derivatives. Study of their reactivity toward phosphine gold(I) chlorides. J. Organomet. Chem., 2010, 695, 2385-2393.10.1016/j.jorganchem.2010.07.019Search in Google Scholar

Brendler E., Heine T., Seichter W., Wagler J., Witter R., 29Si NMR shielding tensors in triphenylsilanes – 29Si solid state NMR experiments and DFT-IGLO calculations. Z. Anorg. Allg. Chem., 2012, 638, 935-944.10.1002/zaac.201200107Search in Google Scholar

Brittain H.G., Polymorphism and solvatomorphism 2010. J. Pharm. Sci., 2012a, 101, 464-484.10.1002/jps.22788Search in Google Scholar PubMed

Brittain H.G., Cocrystal systems of pharmaceutical interest: 2010. Cryst. Growth. Des., 2012b, 12, 1046-1054.10.1021/cg201510nSearch in Google Scholar

Dakternieks D., Altmann R., Jurkschat K., Tiekink E.R.T., Crystal structure of triphenyl[(1,1,1-triphenylstannyl)methyl]stannane, [Ph3SnCH2SnPh]. Z. Kristallogr. NCS, 1998, 213, 525-527.10.1524/ncrs.1998.213.14.553Search in Google Scholar

Dakternieks D., Kuan F.S., Duthie A., Tiekink E.R.T., The crystal structure of the triclinic polymorph of hexaphenyldistannane. Main Group Met. Chem., 2001, 24, 65-66.10.1515/MGMC.2001.24.1.65Search in Google Scholar

Desiraju G.R., The C–H··O hydrogen bond in crystals: what is it? Acc. Chem. Res., 1991, 24, 290-296.10.1021/ar00010a002Search in Google Scholar

Desiraju G.R., Supramolecular synthons in crystal engineering – a new organic synthesis. Angew. Chem. Int. Ed. Engl., 1995, 34, 2311-2327. Angew. Chem., 1995, 107, 2541-2558.10.1002/anie.199523111Search in Google Scholar

Desiraju G.R., Hydrogen bridges in crystal engineering: interactions without borders. Acc. Chem. Res., 2002, 35, 565-573.10.1021/ar010054tSearch in Google Scholar PubMed

Eckardt K., Fuess H., Hattori M., Ikeda R., Ohki H., Weiss A., Structure and dynamics of crystal solvates hexaphenylditin • 2X, X = benzene, toluene, fluorobenzene, chlorobenzene, and aniline. An X-Ray, P(VAPOR)(T) and 2H NMR Study. Z. Naturforsch., 1995, 50a, 758-769.10.1515/zna-1995-0808Search in Google Scholar

Farrugia L.J., WinGX and ORTEP for Windows: an update. J. Appl. Cryst., 2012, 45, 849-854.10.1107/S0021889812029111Search in Google Scholar

Gomberg M., Triphenylmethyl, ein Fall von dreiwerthigem Kohlenstoff. Ber. Dtsch. Chem. Ges., 1900a, 33, 3150-3163 (in German).10.1002/cber.19000330369Search in Google Scholar

Gomberg M., An instance of trivalent carbon: triphenylmethyl. J. Am. Chem. Soc., 1900b, 22, 757-771.10.1021/ja02049a006Search in Google Scholar

Hönle W., Manríquez V., von Schnering H. G., Structures of solvated hexaphenyldisiloxanes. Acta Crystallogr., 1990, C46, 1982-1984.10.1107/S0108270190001640Search in Google Scholar

Karlov S.S., Zaitsev K.V., Lermontova E.Kh., Churakov A.V., CCDC 942103. CSD Communication, 2013.Search in Google Scholar

Krause E., Becker R., Zweiwertiges Zinn als Chromophor in aromatischen Stannoverbindungen und die Gewinnung von Hexaaryl-distannanen. Ber. Dtsch. Chem. Ges., 1920, 53, 173-190 (in German).10.1002/cber.19200530209Search in Google Scholar

Lee A.Y., Erdemir D., Myerson A.S., Crystal polymorphism in chemical process development. Annu. Rev. Chem. Biomol. Eng., 2011, 2, 259-280.10.1146/annurev-chembioeng-061010-114224Search in Google Scholar PubMed

Legon A.C., Millen D.J., Angular geometries and other properties of hydrogen-bonded dimers: a simple electrostatic interpretation of the success of the electron-pair model. Chem. Soc. Rev., 1987, 16, 467-498.10.1039/cs9871600467Search in Google Scholar

Lewars E.G., Hexaphenylethane. In: Lewars E.G. (Ed.), Modeling marvels: computational anticipation of novel molecules. Springer, Dordrecht, 2008; pp 115-129.10.1007/978-1-4020-6973-4_8Search in Google Scholar

Lu F., Cui J., Synthesis and composite crystal structure of a novel organotin complex (C74H60O4Sn4 Synth. React. Inorg. M., 2007, 37, 73-75.10.1080/15533170601187375Search in Google Scholar

Macrae C.F., Edgington P.R., McCabe P., Pidcock E., Shields G.P., Taylor R., et al., Mercury: visualization and analysis of crystal structures. J. Appl. Cryst., 2006, 39, 453-457.10.1107/S002188980600731XSearch in Google Scholar

Nassimbeni L.R., Physicochemical aspects of host-guest compounds. Acc. Chem. Res., 2003, 36, 631-637.10.1021/ar0201153Search in Google Scholar PubMed

Oxford Diffraction Ltd., CrysAlis CCD and CrysAlis RED, Abingdon (U.K.), 2008.Search in Google Scholar

Piana H., Kirchgäßner U., Schubert U., Metall, Wasserstoff, Zinn-Dreizentrenbindung bei Hydrido-Stannyl-Komplexen der 6. Nebengruppe. Chem. Ber., 1991, 124, 743-751 (in German).10.1002/cber.19911240411Search in Google Scholar

Preut H., Haupt H.-J., Huber F., Die Kristall-und Molekularstruktur des Hexaphenyldistannans. Z. Anorg. Allg. Chem., 1973, 396, 81-89 (in German).10.1002/zaac.19733960109Search in Google Scholar

Reeske G., Schürmann M., Jurkschat K., The molecular structure of 1,2-bis(triphenylstannyl)ethane [(C6H53SnCH22 Main Group Met. Chem., 2001, 24, 389-390.10.1515/MGMC.2001.24.6.389Search in Google Scholar

Rexrode N.R., Orien J., King M.D., Effects of solvent stabilization on pharmaceutical crystallization: investigating conformational polymorphism of probucol using combined solid-state density functional theory, molecular dynamics, and terahertz spectroscopy. J. Phys. Chem. A, 2019, 123, 6937-6947.10.1021/acs.jpca.9b00792Search in Google Scholar

Rowland R.S., Taylor R., Intermolecular nonbonded contact distances in organic crystal structures: comparison with distances expected from van der Waals radii. J. Phys. Chem., 1996, 100, 7384-7391.10.1021/jp953141+Search in Google Scholar

Rösel S., Balestrieri B., Schreiner P.R., Sizing the role of London dispersion in the dissociation of all-meta tert-butyl hexaphenylethane. Chem. Sci., 2017, 8, 405-410.10.1039/C6SC02727JSearch in Google Scholar

Schneider-Koglin C., Behrends K., Dräger M., Über gemischte Gruppe 14-Gruppe 14-Bindungen. IV. Hexa-p-tolylethan-Analoga p-Tol6Sn2 p-Tol6PbSn und p-Tol6Pb2 Darstellung, homöotype Strukturen, NMR-Kopplungen und Valenzschwingungen. J. Organomet. Chem., 1991, 415, 349-362 (in German).10.1016/0022-328X(91)80135-7Search in Google Scholar

Schneider-Koglin C., Behrends K., Dräger M., Über gemischte Gruppe 14–Gruppe 14-Bindungen. VI. Hexa-o-tolylethan-Analoga o-Tol6Sn2o-Tol6PbSn und o-Tol6Pb2 ein Vergleich von Bindungsstärke und Polarität in der Reihung Sn–Sn, Pb–Sn, Pb–Pb. J. Organomet. Chem., 1993, 448, 29-38 (in German).10.1016/0022-328X(93)80063-HSearch in Google Scholar

Scotti N., Zachwieja U., Jacobs H., Tetraammin-Lithium-Kationen zur Stabilisierung phenylsubstituierter Zintl-Anionen: Die Verbindung [Li(NH34]2[Sn2Ph4]. Z. Anorg. Allg. Chem., 1997, 623, 1503-1505 (in German).10.1002/zaac.19976231006Search in Google Scholar

Shan N., Zaworotko M.J., The role of cocrystals in pharmaceutical science. Drug Discovery Today, 2008, 13, 440-446.10.1016/j.drudis.2008.03.004Search in Google Scholar PubMed

Sheldrick G.M., SHELXS-97, a program for the solution of crystal structures. Universität Göttingen, Göttingen (Germany), 1997.Search in Google Scholar

Sheldrick G.M., A short history of SHELX. Acta Crystallogr., 2008, A64, 112-122.10.1107/S0108767307043930Search in Google Scholar PubMed

Sheldrick G.M., Crystal structure refinement with SHELXL. Acta Crystallogr., 2015, C71, 3-8.10.1107/S2053229614024218Search in Google Scholar PubMed PubMed Central

Sheldrick G.M., SHELXL-2018. Universität Göttingen, Göttingen (Germany), 2018.Search in Google Scholar

Sita L.R., Bickerstaff R.D., 2,2,4,4,5,5-Hexakis(2,6-diethylphenyl) pentastanna[1.1.1]propellane: characterization and molecular structure. J. Am. Chem. Soc., 1989, 111, 6454-6456.10.1021/ja00198a085Search in Google Scholar

Steed J.W., Atwood J.L., Supramolecular chemistry (2nd ed.). John Wiley & Sons Inc., New York, 2009.10.1002/9780470740880Search in Google Scholar

Steiner T., The hydrogen bond in the solid state. Angew. Chem. Int. Ed., 2002, 41, 48-76. Angew. Chem., 2002, 114, 50-80.10.1002/1521-3773(20020104)41:1<48::AID-ANIE48>3.0.CO;2-USearch in Google Scholar

Turek J., Padelkova Z., Cernosek Z., Erben M., Lycka A., Nechaev M.S., et al., C,N-chelated hexaorganodistannanes, and triorganotin(IV) hydrides and cyclopentadienides. J. Organ omet. Chem., 2009, 694, 3000-3007.10.1016/j.jorganchem.2009.04.043Search in Google Scholar

Wiesler K., Suchentrunk C., Korber N., Syntheses and crystal structures of ammoniates with the phenyl-substituted polytin anions Sn2Ph42−, cyclo-Sn4Ph44−, and Sn6Ph122−. Helv. Chim. Acta., 2006, 89, 1158-1168.10.1002/hlca.200690113Search in Google Scholar

Received: 2019-09-23
Accepted: 2019-11-27
Published Online: 2020-01-31
Published in Print: 2020-01-31

© 2020 Bauer, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/mgmc-2020-0001/html
Scroll to top button