research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Structural parameters of di­methyl sulfoxide, DMSO, at 100 K, based on a redetermination by use of high-quality single-crystal X-ray data

CROSSMARK_Color_square_no_text.svg

aInstitute of Chemistry of New Materials, University of Osnabrück, Barbarastrasse 7, 49069 Osnabrück, Germany
*Correspondence e-mail: hreuter@uos.de

Edited by M. Weil, Vienna University of Technology, Austria (Received 21 July 2017; accepted 29 August 2017; online 5 September 2017)

The title compound, C2H6OS, is a high melting, polar and aprotic solvent widely used in organic and inorganic chemistry. It serves as a H-atom acceptor in hydrogen bonding and is used as an ambidentate ligand in coordination chemistry. The evaluation of the influence of inter­molecular inter­actions on the inter­nal structural parameters of the chemically bonded DMSO mol­ecules affords precise structural data of the free mol­ecule as a point of reference. So far, valid data have been obtained only by use of neutron powder diffraction [Ibberson (2005[Ibberson, R. M. (2005). Acta Cryst. C61, o571-o573.]). Acta Cryst. C61, o571–o573]. In the present redetermination, structural data have been obtained from a single-crystal X-ray diffraction experiment at 100 K, revealing a better comparison with DMSO mol­ecules in other crystal structures. In the solid state, the pyramidal mol­ecule exhibits a nearly perfect Cs symmetry [including H atoms, which are eclipsed with respect to the C⋯C axis], with a C—S—C bond angle of 97.73 (7)° and an S—O bond length of 1.5040 (10) Å, corresponding very well with an S=O double bond, and with almost equal S—C bond lengths [mean value = 1.783 (4) Å] and O—S—C bond angles [mean value = 106.57 (4)°]. The crystal packing is influenced by C—H⋯O inter­actions (2.42–2.47 Å) between all three H atoms of only one methyl group with the O atoms of three neighbouring DMSO mol­ecules. The inter­actions of the O atom with H atoms (or Lewis acids, or hydrogen-donor groups) of adjacent mol­ecules in relation to the orientation of the complete DMSO mol­ecule are described in terms of the angle ω and the distance dnorm; ω is the angle between the pseudo-mirror plane of the mol­ecule and the plane defined through the S=O bond and the inter­acting atom, and dnorm is the distance of the inter­acting atom from the plane perpendicular to the S=O bond.

1. Chemical context

Dimethyl sulfoxide (DMSO), (CH3)2SO, is a colourless polar aprotic solvent with high melting (291 K) and boiling points (462 K), miscible with a wide range of organic solvents and water. It is commonly used in organic and inorganic chemistry because of its capability to dissolve numerous polar or nonpolar compounds. In addition to its solvation properties, the mol­ecule may act as a H-atom acceptor in hydrogen bonding, as well as an ambidentate Lewis base in coordination compounds. In the latter case, DMSO reactivity follows the HSAB principle (Pearson, 1963[Pearson, R. G. (1963). J. Am. Chem. Soc. 85, 3533-3539.]) which means that in combination with `hard' acids like tin(IV), DMSO coordinates via the `hard' O atom [e.g. iPrSnCl3(DMSO-O)2; Kastner & Reuter, 1999[Kastner, G. & Reuter, H. (1999). Main Group Met. Chem. 22, 605-609.]] and in combination with `soft' acids like platinum(II) via the `soft' S atom [e.g. cis-PtCl2(DMSO-S)2; Melanson & Rochon, 1975[Melanson, R. & Rochon, F. D. (1975). Can. J. Chem. 53, 2371-2374.]], while with acids at the `hard–soft' borderline like ruthenium(II), both coordination modes can be realized [cis-RuCl2(DMSO-O)1(DMSO-S)3; Tarighi & Abbasi, 2007[Tarighi, S. & Abbasi, A. (2007). J. Sci. (University of Teheran), 33, 19-21.]]. DMSO is also used in pharmacology in transdermal drug delivery applications and in veterinary medicine.

[Scheme 1]

Both hydrogen-bond formation and formation of coordination bonds will change the structural parameters of the DMSO mol­ecule, as was shown by Calligaris (2004[Calligaris, M. (2004). Coord. Chem. Rev. 248, 351-375.]) for DMSO and other sulfoxides. For the evaluation of the influence of these additional inter­molecular bonds on the inter­nal structural parameters of the coordinating or hydrogen-bonded DMSO ligands, precise data on bond lengths and angles within the free mol­ecule are required as a point of reference. The available data, however, in the case of single-crystal X-ray structure determinations, are from the late 1960s (Viswamitra & Kannan, 1966[Viswamitra, M. A. & Kannan, K. K. (1966). Nature, 209, 1016-1017.]; Thomas et al., 1966[Thomas, R., Shoemaker, C. B. & Eriks, K. (1966). Acta Cryst. 21, 12-20.]) when precession and Weissenberg photographs were state of the art. Therefore, these data are of less accuracy compared with modern X-ray data obtained with CCD area detectors. More recently, Ibberson (2005[Ibberson, R. M. (2005). Acta Cryst. C61, o571-o573.]) published results on neutron powder diffraction studies of fully deuterated dimethyl sulfoxide at 2 and 100 K. Although, the data obtained are of higher precision than those of the forgoing single-crystal X-ray measurements, they suffer from the limitations of powder diffraction techniques.

In the current study, the results of a redetermination of the crystal structure of DMSO based on single-crystal X-ray data at 100 K are presented. The results are comparatively discussed with the previous structure determinations.

2. Structural commentary

Unit-cell parameters of the current 100 K single-crystal X-ray measurement (SCXD) are consistent with those of the neutron powder diffraction (NPD) data of Ibberson (2005[Ibberson, R. M. (2005). Acta Cryst. C61, o571-o573.]), but structural parameters of the DMSO mol­ecule differ considerably between the two refinements (Table 1[link]). In the pyramidal mol­ecule of crystallographic point group symmetry C1 (Fig. 1[link], atom positions and atom labelling according to NPD), the S atom lies 0.6994 (9) Å above the triangular base formed by the O and C atoms. The S—O bond length of 1.5040 (10) Å is slightly longer than the value [1.496 (2) Å] determined by Ibberson at 100 K, but corresponds very well with a S=O double bond in sulfoxides [1.497 (13) Å; Allen et al., 1987[Allen, F. H., Kennard, O., Watson, D. G., Brammer, L., Orpen, A. G. & Taylor, R. (1987). J. Chem. Soc. Perkin Trans. 2, pp. S1-19.]].

Table 1
Experimental details of previous crystal structure determinations of DMSO and their comparison with the present study

  Thomas et al. (1966[Thomas, R., Shoemaker, C. B. & Eriks, K. (1966). Acta Cryst. 21, 12-20.]) Ibberson (2005[Ibberson, R. M. (2005). Acta Cryst. C61, o571-o573.]) This work
Space group, Z P21/c, 4 P21/c, 4 P21/c, 4
a (Å) 5.303 (5) 5.2390 (1) 5.2243 (3)
b (Å) 6.829 (3) 6.7581 (1) 6.7414 (4)
c (Å) 11.693 (3) 11.2696 (1) 11.2772 (6)
β (°) 94.5 (3) 94.8053 (3) 94.820 (2)
V3) 422.2 397.60 (1) 395.77 (4)
T (K) 278 100 100
Sample single-crystal powder single-crystal
Radiation Mo Kα neutron Mo Kα
Technique precession photographs HRPD CDC
R value 7.4% 3.77% 2.4%
Number of reflections 777 not given 938
Number of parameters not given 93 41
H(D) atoms constrained refined constrained
d(S1—O1) (Å) 1.531 (5) 1.496 (2) 1.5040 (10)
d(S1—C1) (Å) 1.775 (8) 1.838 (3) 1.7801 (14)
d(S1—C2) (Å) 1.821 (11) 1.788 (3) 1.7861 (15)
O1—S1—C1 (°) 106.7 (4) 105.2 (2) 106.54 (6)
O1—S1—C2 (°) 106.8 (4) 108.3 (2) 106.60 (7)
C1—S1—C2 (°) 97.4 (4) 96.4 (1) 97.73 (7)
[Figure 1]
Figure 1
The molecular structure of the title compound, showing the atom-labelling scheme and displacement ellipsoids for the non-H atoms at the 50% probability level.

Other differences between the single-crystal X-ray and neutron powder diffraction data, however, are strongly expressed with respect to S—C bond lengths and even more with respect to O—S—C bond angles (Table 1[link]). In the case of the neutron data, the difference between both S—C bonds is 0.05 Å [S—C1 = 1.838 (3) Å and S—C2 = 1.788 (3) Å], while in the case of the X-ray data, the difference between both bonds is reduced by a factor of about 10 to 0.006 Å [S—C1 = 1.7801 (14) Å and S—C2 = 1.7861 (15) Å]. Moreover, the bond to atom C1 is shorter than the bond to C2, in contrast to the bond-length distribution observed by Ibberson. This is of special inter­est in view of the C—H⋯O inter­actions discussed below. With respect to the C—S—O bond angles, structural differences between the NPD and SCXD model are enormous: the difference between both bond angles of 3.09° [O—S—C1 = 105.21 (16)° and O—S—C2 = 108.30 (15)°] found by Ibberson at 100 K can be compared with a difference of only 0.06° [O—S—C1 = 106.54 (6)° and O—S—C2 = 106.60 (7)°] in the case of the present work. All in all, the ideal Cs point group symmetry of the gaseous and liquid DMSO mol­ecule is much better approached in the crystalline state, even at 100 K, than originally assumed from neutron powder data.

Although this symmetry consideration is not affected by the bond angle between the S atom and the methyl groups, it is important – on the background of coordination that seems to have a great influence on this bond angle – to emphasize that in the SCXD model [C—S—C = 97.73 (3)°], this angle is about 1.4° larger than in the NPD model [C—S—C = 96.37 (12)° at 100 K]. With respect to the hydrogen/deuterium positions, no differences occur, as both methyl groups show an eclipsed orientation with respect to C⋯C, thus fulfilling the nearly ideal Cs symmetry, too.

For the sake of completeness, structural data of the previous single-crystal X-ray structure determination by Thomas et al. (1966[Thomas, R., Shoemaker, C. B. & Eriks, K. (1966). Acta Cryst. 21, 12-20.]) are also compiled in Table 1[link].

3. Supra­molecular features

C—H⋯O contacts are the most prominent inter­molecular inter­actions responsible for the three-dimensional arrangement of the DMSO mol­ecules in the solid state (Fig. 2[link]). In order to compare our results with the results of the neutron powder diffraction experiment, one must take into account the different validity and refinement strategies for the H/D atoms in both methods. Under consideration of the van der Waals radii of H (1.10 Å) and O (1.52 Å) supplied by Mantina et al. (2009[Mantina, M., Chamberlin, A. C., Valero, R., Cramer, C. J. & Truhlar, D. G. (2009). J. Phys. Chem. A, 113, 5806-5812.]), relevant H⋯O distances should be shorter than 2.62 Å. From the H atoms attached to C2, only one (H22) shows an inter­atomic distance below this threshold. With an H22⋯O1ii (for symmetry code, see Table 2[link]) distance of 2.61 Å, a binding C—H⋯O inter­action other than a van der Waals inter­action can be excluded. Just the opposite is observed in case of the H atoms attached to C1: all three H atoms show an inter­molecular contact to one O atom of three different DMSO mol­ecules in the range 2.42–2.47 Å (Table 2[link]). In this case, these contacts fall below the van der Waals distance by 7.6–5.7% (0.20–0.15 Å) which justifies the assumption of binding C—H⋯O inter­actions. The corresponding inter­molecular donor–acceptor distances are in the range 3.318 (2)–3.445 (2) Å, while the C—H⋯O angles are in the range 152.0–173.0° (Table 2[link]). In summary, each DMSO mol­ecule participates in six C—H⋯O contacts to five neighbouring mol­ecules (Fig. 3[link]). The extent of the van der Waals and hydrogen-bonding inter­actions on the overlapping of the mol­ecules is visualized in Fig. 4[link]. Obviously, there is no weakening influence of these inter­actions on the S—C bond length. Quite the opposite, the S1—C1 bond is somewhat shorter than the S1—C2 bond (see above).

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
C1—H11⋯O1i 0.98 2.42 3.3318 (18) 155
C1—H12⋯O1ii 0.98 2.42 3.3184 (17) 152
C1—H13⋯O1iii 0.98 2.47 3.4450 (19) 173
C2—H22⋯O1ii 0.98 2.61 3.4618 (18) 146
Symmetry codes: (i) [-x, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (ii) x+1, y, z; (iii) -x, -y, -z.
[Figure 2]
Figure 2
Crystal packing showing the tube-like arrangment of the mol­ecules along the b axis.
[Figure 3]
Figure 3
C—H⋯O contacts (grey brocken sticks) between the DMSO mol­ecule and its neighbours. [Symmetry codes used to generate equivalent atoms: (i) 1 − x, [{1\over 2}] + y, [{1\over 2}] − z; (ii) −1 + x, y, z; (iii) −x, −y, −z; (iv) 1 + x, y, z; (v) −x, −[{1\over 2}] + y, [{1\over 2}] − z.]
[Figure 4]
Figure 4
Space-filling model of the DMSO mol­ecule showing four of the six C—H⋯O contacts. Spheres overlap has been visualized by removing the resulting caps.

With respect to the O atom as an acceptor atom, bond angles (S=O⋯H) of the van der Waals contacts come to 114.0° for H132 (2 = −1 + x, y, z), 152.1° for H121 (1 = −x, 0.5 + y, 0.5 − z), and 103.4° for H113 (3 = −x, −y, −z). The geometrical aspects of these van der Waals inter­actions (or of coordinatively or hydrogen-bonded DMSO mol­ecules) are described only incompletely with the foregoing used distances and angles as they disregard the orientation of the complete DMSO mol­ecule in relation to the inter­actions described. In order to unambiguously account for this specific relationship, indexation by two additional values, ω and dnorm, using two planes as a reference (Fig. 5[link]) is suggested. The first plane is identical, with the pseudo-mirror plane m′ defined by O1, S1 and the mid-point between both C atoms. The second plane, plO, is perpendicular to the S=O bond and located in O1. While dnorm represents the distance between the inter­acting atom (via a van der Waals inter­action, a hydrogen bond or a coordinative bond) and plO, the angle ω marks the angle between m′ and the plane plH defined by O1, S1 and the inter­acting atom. Values of ω can stretch from 0 to 360° when looking down the O=S bond as in a Newman projection. In the case of the van der Waals inter­actions discussed here, the corresponding ω/dnorm values are: H121 = 98.2°/2.139 Å, H132 = 178.3°/1.006 Å and H113 = 335.1°/0.560 Å.

[Figure 5]
Figure 5
Geometrical boundary conditions for the determination of ω (left) and dnorm (right, distances in Å) by use of the pseudo-mirror plane m′ (definition: O1, S1, mid-point between C1 and C2), the plane plH (definition: the inter­acting atom, S=O bond), and the plane plO (definition: O1, S=O bond = normal vector). [Symmetry codes used to generate equivalent atoms: (i) 1 − x, [{1\over 2}] + y, [{1\over 2}] − z; (ii) −1 + x, y, z; (iii) −x, −y, −z.]

4. Synthesis and crystallization

Single crystals were grown from a commercial available sample (Sigma–Aldrich) within a 0.3 mm thick Lindemann capillary using the Kryoflex low-temperature device of the diffractometer.

5. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. All six H atoms were found in a difference-Fourier map. They could be refined without any restraints in meaningful positions [C—H range = 0.91 (2)–0.97 (2) Å; H—C—H range = 107.6 (11)–112.2 (15)°] with individual isotropic displacement parameters [range = 0.018 (4)–0.036 (5) Å2]. In order to obtain a structure model comparable to typical refinement techniques of DMSO mol­ecules in the structures of coordination compounds or with hydrogen bonds, conventional constraints [AFIX 137, C—H = 0.99 Å, H—C—H = 109.6° in SHELXL (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.])] with two common isotropic displacement parameters, one for each methyl group, have been applied. In summary, these restraints only slightly affected the final results: the final R value increased from 2.37 to 2.42%, while the bond lengths and bond angles remained unchanged. All data have been approved by a second independently grown crystal.

Table 3
Experimental details

Crystal data
Chemical formula C2H6OS
Mr 78.13
Crystal system, space group Monoclinic, P21/c
Temperature (K) 100
a, b, c (Å) 5.2243 (3), 6.7414 (4), 11.2772 (6)
β (°) 94.820 (2)
V3) 395.77 (4)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.60
Crystal size (mm) 0.21 × 0.17 × 0.16
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Bruker, 2009[Bruker (2009). APEX2, SADABS, SAINT and SHELXTL. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.883, 0.913
No. of measured, independent and observed [I > 2σ(I)] reflections 5476, 938, 801
Rint 0.030
(sin θ/λ)max−1) 0.659
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.024, 0.062, 1.13
No. of reflections 938
No. of parameters 41
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.26, −0.25
Computer programs: APEX2 (Bruker, 2009[Bruker (2009). APEX2, SADABS, SAINT and SHELXTL. Bruker AXS Inc., Madison, Wisconsin, USA.]), SAINT (Bruker, 2009[Bruker (2009). APEX2, SADABS, SAINT and SHELXTL. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2014 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), DIAMOND (Brandenburg, 2006[Brandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.]), Mercury (Macrae et al. (2008[Macrae, C. F., Bruno, I. J., Chisholm, J. A., Edgington, P. R., McCabe, P., Pidcock, E., Rodriguez-Monge, L., Taylor, R., van de Streek, J. & Wood, P. A. (2008). J. Appl. Cryst. 41, 466-470.]) and SHELXTL (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]).

Supporting information


Computing details top

Data collection: APEX2 (Bruker, 2009); cell refinement: SAINT (Bruker, 2009); data reduction: SAINT (Bruker, 2009); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008); program(s) used to refine structure: SHELXL2014 (Sheldrick, 2015); molecular graphics: DIAMOND (Brandenburg, 2006) and Mercury (Macrae et al. (2008); software used to prepare material for publication: SHELXTL (Sheldrick, 2008).

Dimethyl sulfoxide top
Crystal data top
C2H6OSF(000) = 168
Mr = 78.13Dx = 1.311 Mg m3
Monoclinic, P21/cMo Kα radiation, λ = 0.71073 Å
a = 5.2243 (3) ÅCell parameters from 3349 reflections
b = 6.7414 (4) Åθ = 3.5–28.0°
c = 11.2772 (6) ŵ = 0.60 mm1
β = 94.820 (2)°T = 100 K
V = 395.77 (4) Å3Bloc, colourless
Z = 40.21 × 0.17 × 0.16 mm
Data collection top
Bruker APEXII CCD
diffractometer
801 reflections with I > 2σ(I)
φ and ω scansRint = 0.030
Absorption correction: multi-scan
(SADABS; Bruker, 2009)
θmax = 28.0°, θmin = 3.5°
Tmin = 0.883, Tmax = 0.913h = 66
5476 measured reflectionsk = 88
938 independent reflectionsl = 1414
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.024H-atom parameters constrained
wR(F2) = 0.062 w = 1/[σ2(Fo2) + (0.0238P)2 + 0.1531P]
where P = (Fo2 + 2Fc2)/3
S = 1.13(Δ/σ)max < 0.001
938 reflectionsΔρmax = 0.26 e Å3
41 parametersΔρmin = 0.25 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
S10.18232 (6)0.14961 (5)0.19283 (3)0.01776 (12)
O10.10440 (18)0.13580 (16)0.16758 (10)0.0245 (3)
C10.3139 (3)0.0633 (2)0.12708 (13)0.0187 (3)
H110.26630.18250.16980.027 (3)*
H120.50150.05180.13170.027 (3)*
H130.24650.07290.04350.027 (3)*
C20.2905 (3)0.3307 (2)0.09280 (15)0.0247 (3)
H210.22170.29940.01140.033 (3)*
H220.47860.32990.09720.033 (3)*
H230.23050.46230.11490.033 (3)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
S10.01277 (18)0.02359 (19)0.0170 (2)0.00180 (13)0.00161 (12)0.00362 (15)
O10.0115 (5)0.0353 (6)0.0269 (6)0.0008 (4)0.0030 (4)0.0069 (5)
C10.0159 (6)0.0194 (7)0.0209 (8)0.0003 (5)0.0030 (6)0.0000 (6)
C20.0236 (7)0.0198 (7)0.0312 (9)0.0001 (6)0.0042 (6)0.0018 (6)
Geometric parameters (Å, º) top
S1—O11.5040 (10)C1—H130.9800
S1—C11.7801 (14)C2—H210.9800
S1—C21.7861 (15)C2—H220.9800
C1—H110.9800C2—H230.9800
C1—H120.9800
O1—S1—C1106.54 (6)H12—C1—H13109.5
O1—S1—C2106.60 (7)S1—C2—H21109.5
C1—S1—C297.73 (7)S1—C2—H22109.5
S1—C1—H11109.5H21—C2—H22109.5
S1—C1—H12109.5S1—C2—H23109.5
H11—C1—H12109.5H21—C2—H23109.5
S1—C1—H13109.5H22—C2—H23109.5
H11—C1—H13109.5
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
C1—H11···O1i0.982.423.3318 (18)155
C1—H12···O1ii0.982.423.3184 (17)152
C1—H13···O1iii0.982.473.4450 (19)173
C2—H22···O1ii0.982.613.4618 (18)146
Symmetry codes: (i) x, y1/2, z+1/2; (ii) x+1, y, z; (iii) x, y, z.
Experimental details of previous crystal structure determinations of DMSO and their comparison with the present study. top
Thomas et al. (1966)Ibberson (2005)This work
Space group, ZP21/c, 4P21/c, 4P21/c, 4
a (Å)5.303 (5)5.2390 (1)5.2243 (3)
b (Å)6.829 (3)6.7581 (1)6.7414 (4)
c (Å)11.693 (3)11.2696 (1)11.2772 (6)
β (°)94.5 (3)94.8053 (3)94.820 (2)
V3)422.2397.60 (1)395.77 (4)
T (K)278100100
Samplesingle crystalpowdersingle crystal
RadiationMo KαneutronMo Kα
Techniqueprecession photographsHRPDCCDC
R value7.4%3.77%2.4%
Number of reflections777not given938
Number of parametersnot given9341
H(D) atomsconstrainedrefinedconstrained
d(S1—O1) (Å)1.531 (5)1.496 (2)1.5040 (10)
d(S1—C1) (Å)1.775 (8)1.838 (3)1.7801 (14)
d(S1—C2) (Å)1.821 (11)1.788 (3)1.7861 (15)
O1—S1—C1 (°)106.7 (4)105.2 (2)106.54 (6)
O1—S1—C2 (°)106.8 (4)108.3 (2)106.60 (7)
C1—S1—C2 (°)97.4 (4)96.4 (1)97.73 (7)
 

Acknowledgements

We thank the Deutsche Forschungsgemeinschaft and the Government of Lower-Saxony for funding the diffractometer and acknowledge support by Deutsche Forschungsgemeinschaft (DFG) and Open Access Publishing Fund of Osnabrück University.

References

First citationAllen, F. H., Kennard, O., Watson, D. G., Brammer, L., Orpen, A. G. & Taylor, R. (1987). J. Chem. Soc. Perkin Trans. 2, pp. S1–19.  CSD CrossRef Web of Science
First citationBrandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.
First citationBruker (2009). APEX2, SADABS, SAINT and SHELXTL. Bruker AXS Inc., Madison, Wisconsin, USA.
First citationCalligaris, M. (2004). Coord. Chem. Rev. 248, 351–375.  Web of Science CrossRef CAS
First citationIbberson, R. M. (2005). Acta Cryst. C61, o571–o573.  Web of Science CSD CrossRef CAS IUCr Journals
First citationKastner, G. & Reuter, H. (1999). Main Group Met. Chem. 22, 605–609.  CrossRef CAS
First citationMacrae, C. F., Bruno, I. J., Chisholm, J. A., Edgington, P. R., McCabe, P., Pidcock, E., Rodriguez-Monge, L., Taylor, R., van de Streek, J. & Wood, P. A. (2008). J. Appl. Cryst. 41, 466–470.  Web of Science CSD CrossRef CAS IUCr Journals
First citationMantina, M., Chamberlin, A. C., Valero, R., Cramer, C. J. & Truhlar, D. G. (2009). J. Phys. Chem. A, 113, 5806–5812.  Web of Science CrossRef PubMed CAS
First citationMelanson, R. & Rochon, F. D. (1975). Can. J. Chem. 53, 2371–2374.  CSD CrossRef CAS Web of Science
First citationPearson, R. G. (1963). J. Am. Chem. Soc. 85, 3533–3539.  CrossRef CAS Web of Science
First citationSheldrick, G. M. (2008). Acta Cryst. A64, 112–122.  Web of Science CrossRef CAS IUCr Journals
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals
First citationTarighi, S. & Abbasi, A. (2007). J. Sci. (University of Teheran), 33, 19–21.  CAS
First citationThomas, R., Shoemaker, C. B. & Eriks, K. (1966). Acta Cryst. 21, 12–20.  CSD CrossRef IUCr Journals Web of Science
First citationViswamitra, M. A. & Kannan, K. K. (1966). Nature, 209, 1016–1017.  CSD CrossRef CAS Web of Science

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds