Skip to content
BY 4.0 license Open Access Published by De Gruyter (O) September 18, 2019

Crystal structure of catena-poly[tri(4-chlorophenyl)-(μ2-hydroxido)tin(IV)] – 2-propanol (1/1), C21H21Cl3O2Sn

  • Kong Mun Lo , See Mun Lee and Edward R.T. Tiekink ORCID logo EMAIL logo

Abstract

C21H21Cl3O2Sn, monoclinic, P21/n (no. 14), a = 13.3360(1) Å, b = 8.2332(1) Å, c = 20.2443(2) Å, β = 104.070(1)°, V = 2156.10(4) Å3, Z = 4, Rgt(F) = 0.0199, wRref(F2) = 0.0544, T = 100(2) K.

CCDC no.: 1951436

Table 1 contains crystallographic data and Table 2 contains the list of the atoms including atomic coordinates and displacement parameters.

Table 1:

Data collection and handling.

Crystal:Colourless prism
Size:0.16 × 0.07 × 0.03 mm
Wavelength:Cu Kα radiation (1.54178 Å)
μ:13.0 mm−1
Diffractometer, scan mode:XtaLAB Synergy, ω
θmax, completeness:67.1°, >99%
N(hkl)measured, N(hkl)unique, Rint:49159, 3845, 0.036
Criterion for Iobs, N(hkl)gt:Iobs > 2 σ(Iobs), 3647
N(param)refined:252
Programs:CrysAlisPRO [1], SHELX [2], [3], WinGX/ORTEP [4]
Table 2:

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2).

AtomxyzUiso*/Ueq
Sn0.74641(2)0.62820(2)0.25099(2)0.01366(6)
Cl10.85701(4)0.45614(6)0.58296(2)0.02580(11)
Cl21.14436(5)0.86190(7)0.12846(4)0.04162(15)
Cl30.29619(4)0.47006(6)0.03131(2)0.03108(12)
O10.70141(10)0.87677(13)0.26734(7)0.0165(3)
H1O0.6489(13)0.881(3)0.2829(12)0.025*
C10.78181(13)0.5769(2)0.35770(9)0.0165(3)
C20.72610(15)0.6470(2)0.40081(10)0.0202(4)
H20.67200.72140.38260.024*
C30.74872(16)0.6094(2)0.46977(10)0.0222(4)
H30.71010.65680.49850.027*
C40.82802(14)0.5023(2)0.49600(9)0.0201(4)
C50.88599(14)0.4329(2)0.45525(9)0.0201(4)
H50.94100.36060.47410.024*
C60.86223(13)0.4710(2)0.38622(9)0.0181(4)
H60.90170.42370.35790.022*
C70.87130(13)0.7107(2)0.21065(9)0.0158(3)
C80.96992(14)0.7304(2)0.25384(9)0.0217(4)
H80.97970.71150.30130.026*
C91.05356(15)0.7768(2)0.22895(10)0.0270(4)
H91.12000.78970.25880.032*
C101.03870(16)0.8041(2)0.15984(11)0.0268(4)
C110.94273(16)0.7860(2)0.11552(10)0.0279(4)
H110.93370.80560.06820.033*
C120.85949(14)0.7388(2)0.14113(9)0.0217(4)
H120.79340.72550.11080.026*
C130.60105(13)0.5794(2)0.18139(9)0.0171(3)
C140.50882(15)0.6493(2)0.18890(10)0.0206(4)
H140.51030.72140.22580.025*
C150.41476(16)0.6150(2)0.14327(11)0.0232(4)
H150.35240.66190.14910.028*
C160.41373(14)0.5117(2)0.08945(9)0.0222(4)
C170.50345(15)0.4420(2)0.08004(9)0.0222(4)
H170.50150.37140.04260.027*
C180.59676(14)0.4767(2)0.12617(9)0.0196(4)
H180.65870.42940.11980.024*
O20.52056(12)0.88333(16)0.32963(9)0.0302(3)
H2O0.507(2)0.9835(14)0.3316(14)0.045*
C190.44785(16)0.7908(3)0.35541(11)0.0300(4)
H190.46200.67330.34900.036*
C200.4606(2)0.8188(5)0.43005(13)0.0641(9)
H20A0.53180.79360.45440.096*
H20B0.41290.74840.44670.096*
H20C0.44550.93260.43790.096*
C210.33907(19)0.8250(4)0.31621(14)0.0518(7)
H21A0.32170.93830.32340.078*
H21B0.29140.75270.33200.078*
H21C0.33310.80650.26760.078*

Source of material

All chemicals and solvents were used as purchased without purification. The melting point was determined using a Mel-temp II digital melting point apparatus and was uncorrected. The IR spectrum was obtained on a Perkin-Elmer RX1 spectrophotometer in a Nujol mull between KBr plates from 4000 to 400 cm−1.

Tetra(4-chlorophenyl)tin was prepared from the reaction of stannic chloride (Fluka) and 4-chlorophenylmagnesium bromide (Fluka) in a 1:4 molar ratio. Tetra(4-chlorophenyl)tin (3.40 g, 6 mmol) was dissolved in dimethylformamide (30 mL) and cooled to 273 K in an ice-salt bath. A chilled dimethylformamide solution (5 mL) of bromine (0.97 g, 6 mmol) was added drop-wise into the solution. When the addition was complete, the mixture was stirred at room temperature for 3 h and a 10% sodium hydroxide solution was added drop-wise until a white precipitate was formed. The solid was filtered and recrystallised from 2-propanol to afford the title compound as colourless crystals. Yield: 0.30 g (10%). M.pt: 431–433 K. IR (cm−1) 1568 (m) ν(C—C), 1090 (s) ν(C—O), 1010 (s) ν(C—O), 481 (w) ν(Sn—O).

Experimental details

The C-bound H atoms were geometrically placed (C—H = 0.95–1.00 Å) and refined as riding with Uiso(H) = 1.2–1.5Ueq(C). The O-bound H-atoms were located in a difference Fourier map but were refined with a distance restraint of O—H = 0.84 ± 0.01 Å, and with Uiso(H) set to 1.5Ueq(O).

Comment

Molecules of the general formula R3SnOH can be mononuclear as exemplified by R = 2,4,6-Me3C6H2 (mesityl) [5], binuclear, as in the case when R = CF2CF3 [6], or polymeric such as when R = Et [7]. While to a first approximation, the adoption of one structural motif over another for R3SnOH may be attributed to steric effects, well established in organotin chemistry to moderate secondary bonding interactions [8], [9], [10], [11], the subtle nature of the adoption of one motif over another is nicely illustrated in the case when R = Ph. Thus, the structure of Ph3SnOH was established long ago as being polymeric [12] but, when this is co-crystallised with 2,6-bis(1H-benzimidazol-2-yl)pyridine and water [13], it is mononuclear. Herein, the crystal and molecular structures of the R = 4-ClPh derivative, isolated as a 1:1 isopropanol solvate, that is (4-ClPh)3Sn(OH)⋅iPrOH, (I), are described.

The crystallographic asymmetric unit of (I) is shown in the upper part of the figure (70% probability displacement ellipsoids). From symmetry, and as indicated in the lower view of the figure (only ipso-C atoms of the 4-ClPh substituents are shown and the solvent molecules are omitted), the solid-state structure of (I) is a one-dimensional coordination polymer with helical symmetry, being propagated along the b-axis by 21-screw symmetry. The tin atom is therefore coordinated by two hydroxy-O atoms and three ipso-C atoms of the organic substiuents. While, the Sn—C bond lengths are experimentally equivalent, that is, [2.1363(17) Å, for Sn—C7, to 2.1393(17) Å, for Sn—C13], there is a disparity in the Sn—O1 bond lengths with the formally covalent Sn—O1 bond length of 2.1805(11) Å being significantly shorter than the formally intermolecular Sn—O1i bond length of 2.2434(11) Å; symmetry operation (i): 3/2 − x, −1/2 + y, 1/2 − z. As the O1 atoms are approximately trans to each other, the O1—Sn—O1i bond angle is 177.51(2)°, the resultant C3O2 coordination geometry is close to trigonal-bipyramidal.

As indicated in the upper view of the figure, the connection between the constituents of the asymmetric unit is a tin-hydroxy-O—H⋯O(2-propanol) hydrogen bond [O1—H1o⋯O2: H1o⋯O2 = 2.148(19) Å, O1⋯O2 = 2.981(2) Å with angle at H1o = 176(2)°]. The acidic proton of the solvent iPrOH molecule is engaged in an intra-chain 2-propanol-O—H⋯π(phenyl) interaction [O2—H2o⋯π(C7—C12)ii: H2o⋯Cg(C7—C12)ii = 2.353(13) Å, O2⋯Cg(C7—C12)ii = 3.1203(15) Å with angle at H2o = 151(2)° for symmetry operation (ii): 3/2 − x, 1/2 + y, 1/2 − z]. Chains pack in the crystal with no directional interactions between them.

To analyse the molecular packing further, using Crystal Explorer 17 [14] and established protocols [15], Hirshfeld surface calculations were performed along with the full and delineated two-dimensional fingerprint plots. This analysis reveals the importance of Cl⋯H/H⋯Cl (32.2%), H⋯H (30.2%) and C⋯H/H⋯C (25.6%) contacts to the calculated Hirshfeld surface but, these occur at distances generally greater than the sum of the respective van der Waals radii. The next most important contacts are of the type C⋯C, contributing 3.8%.

As indicated above, there are several polymeric precedents for the one-dimensional coordination polymer found in (I). Zig-zag chains are found in the R = Et derivative [7] and in the mixed species, Ph2(Cy)SnOH [16], each of which sees the chain having mirror symmetry. Helical chains propagated by 21 scew axes are found when R = Ph [12], R = CH2Ph [17] and in the mixed species Me2(PhCH2)SnOH [18]. The discussion of these structures would be incomplete without mentioning the structural analyses of the R = Me derivative, which has attracted considerable attention [19]. Since the original room temperature analysis where a helical chain was shown [20], in a subsequent study, {Me3Sn(OH)}n was modelled as a 32-fold superstructure with 83-helical chains [21]. At 220 K, the structure has been described as an eight-fold superstructure, with distorted zig-zag chains [22].

Funding source: Sunway University Sdn Bhd is thanked for financial support

Award Identifier / Grant number: STR-RCTR-RCCM-001-2019

Funding statement: Sunway University Sdn Bhd is thanked for financial support of this work through Grant no. STR-RCTR-RCCM-001-2019.

References

1. Agilent Technologies: CrysAlisPRO. Agilent Technologies, Santa Clara, CA, USA (2017).Search in Google Scholar

2. Sheldrick, G. M.: A short history of SHELX. Acta Crystallogr. A64 (2008) 112–122.10.1107/S0108767307043930Search in Google Scholar

3. Sheldrick, G. M.: Crystal structure refinement with SHELXL. Acta Crystallogr. C71 (2015) 3–8.10.1107/S2053229614024218Search in Google Scholar

4. Farrugia, L. J.: WinGX and ORTEP for Windows: an update. J. Appl. Crystallogr. 45 (2012) 849–854.10.1107/S0021889812029111Search in Google Scholar

5. Reuter, H.; Puff, H.: Trimesitylzinn-hydroxid und -fluorid als modellsubstanzen zur Bestimmung der normal-Kovalenzabstände Zinn-Sauerstoff und Zinn-Fluor. J. Organomet. Chem. 379 (1989) 223–234.10.1016/0022-328X(89)85162-9Search in Google Scholar

6. Wiesemann, M.; Hoge, B.: Pentafluoroethylated compounds of silicon, germanium and tin. Chem. Eur. J. 24 (2018) 16457–16471.10.1002/chem.201801292Search in Google Scholar PubMed

7. Deacon, G. B.; Lawrenz, E.; Nelson, K. T.; Tiekink, E. R. T.: Crystal structure of polymeric triethyltin hydroxide. Main Group Met. Chem. 16 (1993) 265–269.Search in Google Scholar

8. Tiekink, E. R. T.: Structural chemistry of organotin carboxylates: a review of the crystallographic literature. Appl. Organomet. Chem. 5 (1991) 1–23.10.1002/aoc.590050102Search in Google Scholar

9. Buntine, M. A.; Kosovel, F. J.; Tiekink, E. R. T.: Supramolecular Sn⋯Cl associations in diorganotin dichlorides and their influence on molecular geometry as studied by ab initio molecular orbital calculations. CrystEngComm 5 (2003) 331–336.10.1039/B308922CSearch in Google Scholar

10. Tiekink, E. R. T.: Tin dithiocarbamates: applications and structures. Appl. Organomet. Chem. 22 (2008) 533–550.10.1002/aoc.1441Search in Google Scholar

11. Tiekink, E. R. T.: Supramolecular assembly based on “emerging” intermolecular interactions of particular interest to coordination chemists. Coord. Chem. Rev. 345 (2017) 209–228.10.1016/j.ccr.2017.01.009Search in Google Scholar

12. Glidewell, C.; Liles, D. C.: The crystal and molecular structures of hydroxotriphenyltin(IV) and hydroxotriphenyllead(IV). Acta Crystallogr. B34 (1978) 129–134.10.1107/S0567740878002447Search in Google Scholar

13. Lo, K. M.; Das, V. G. K.; Ng, S. W.: Hydroxotriphenyltin 2,6-bis(1H-benzimidazol-2-yl)pyridine hydrate. Acta Crystallogr. C55 (1999) 1234–1236.10.1107/S0108270199004072Search in Google Scholar

14. Turner, M. J.; Mckinnon, J. J.; Wolff, S. K.; Grimwood, D. J.; Spackman, P. R.; Jayatilaka, D.; Spackman, M. A.: Crystal Explorer v17. The University of Western Australia, Australia (2017).Search in Google Scholar

15. Tan, S. L.; Jotani, M. M.; Tiekink, E. R. T.: Utilizing Hirshfeld surface calculations, non-covalent interaction (NCI) plots and the calculation of interaction energies in the analysis of molecular packing. Acta Crystallogr. E75 (2019) 308–318.10.1107/S2056989019001129Search in Google Scholar

16. Lo, K. M.; Ng, S. W.: catena-Poly[[cyclohexyldiphenyltin(IV)]-μ-hydroxido-κ2O:O]. Acta Crystallogr. E64 (2008) m724–m725.10.1107/S1600536808011100Search in Google Scholar

17. Reuter, H.: Crystal structure of tribenzyltin(IV) hydroxide, (C7H7)3SnOH. Z. Kristallogr. NCS 219 (2004) 487–488.10.1524/ncrs.2004.219.14.519Search in Google Scholar

18. Wannagat, U.; Damrath, V.; Huch, V.; Veith, M.; Harder, U.: Sila-Riechstoffe und Riechstoffisostere XII. Geruchsvergleiche homologer organoelementverbindungen der vierten Hauptgruppe (C, Si, Ge, Sn). J. Organomet. Chem. 443 (1993) 153–165.10.1016/0022-328X(93)80293-KSearch in Google Scholar

19. Dey, S.; Schonleber, A.; Mondal, S.; van Smaalen, S.: Superspace description of trimethyltin hydroxide at T = 100 K. Z. Kristallogr. – Cryst. Mater. 231 (2016) 427–434.10.1515/zkri-2016-1952Search in Google Scholar

20. Kasai, N.; Yasuda, K.; Okawara, R.: The crystal structure of trimethyltin hydroxide. J. Organomet. Chem. 3 (1965) 172–173.10.1016/S0022-328X(00)84752-XSearch in Google Scholar

21. Anderson, K. M.; Tallentire, S. E.; Probert, M. R.; Goeta, A. E.; Mendis, B. G.; Steed, J. W.: Trimethyltin hydroxide: a crystallographic and high Z′ curiosity. Cryst. Growth Des. 11 (2011) 820–826.10.1021/cg101464jSearch in Google Scholar

22. Dey, S.; Schonleber, A.; Mondal, S.; Ali, S. I.; van Smaalen, S.: Role of steric hindrance in the crystal packing of Z′ = 4 superstructure of trimethyltin hydroxide. Cryst. Growth Des. 18 (2018) 1394–1400.10.1021/acs.cgd.7b01295Search in Google Scholar

Received: 2019-08-16
Accepted: 2019-09-04
Published Online: 2019-09-18
Published in Print: 2019-12-18

©2019 Kong Mun Lo et al., published by De Gruyter, Berlin/Boston

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

Downloaded on 29.3.2024 from https://www.degruyter.com/document/doi/10.1515/ncrs-2019-0552/html
Scroll to top button