Abstract
Complexes of Sn(II), Sn(IV), Ge(IV), and Si(IV) with the ambidentate pyridine-2-thiolato ligand (PyS-) were synthesized and characterized by multinuclear NMR spectroscopy and single-crystal X-ray diffractometry. Comparison of the structures of E(PyS)2Cl2 (E=Sn, Ge, Si) and E(PyS)4 (E=Sn, Si) allows for insights into the group 14 coordination chemistry of this ambidentate chelator in dependence of the thiophilicity of the central atom of the corresponding complex. Furthermore, the crystal structure of Sn(PyS)2 reveals two different coordination modes of its constituents, i.e., the crystal packing features cyclic dimers and polymeric chains of Sn(PyS)2. This compound was shown to undergo oxidative addition of 2,2′-dipyridyldisulfide and sulfur with the formation of Sn(PyS)4and (PyS)2Sn(μ-S)2Sn(PyS)2, respectively.
Introduction
The anion of methimazole (2-mercapto-1-methylimidazole, 1-methylimidazoline-2-thione, Hmt) proved to be a suitable buttress to link soft transition metal atoms (from group 10) with hard group 14 element atoms, thus supporting bonds between the group 10 and group 14 elements (e.g., Scheme 1, compounds I and II) (Wagler et al., 2008, 2010a; Brendler et al., 2011; Truflandier et al., 2011; Autschbach et al., 2012). The ambidentate nature of the methimazolide (mt-) ligand supports the formation of heteronuclear complexes that bear this buttress S-bound to the group 10 metal and N-bound to the group 14 element. In a similar manner, this ligand was shown to buttress heteronuclear complexes of different group 14 elements (Scheme 1, III) (Wagler et al., 2010b), whereas in case of related homo-dinuclear complexes, a so-called (2,2)-paddlewheel arrangement is favored (Scheme 1, IV) (Wagler et al., 2013).

mt-bridged hetero- and homo-dinuclear group 14 complexes.
Recently, the anion of 2-mercaptopyridine (better called the anion of 1,2-dihydropyridine-2-thione, pyridine-2-thiolate, PyS-) has been successfully utilized as a related (S,N) ambidentate buttress for the syntheses of heteronuclear complexes of group 10 and group 14 elements (Scheme 2, V and VI) (Martincová et al., 2011, 2012; Berrenguer et al., 2013). This tempted us to explore the chemistry of this buttressing ligand in systems related to those with methimazolyl buttresses. In terms of group 14 element coordination chemistry, to date, the Cambridge Structural Database contains examples of only Sn and Pb compounds of pyridine-2-thiolate; Ge and Si compounds are lacking (ConQuest 1.15, 2012). In tin(IV) chemistry, PyS- and related ligands have been shown capable of enhancing the tin coordination number to 5, 6, 7, and even 8 (Scheme 3, VII, VIII and IX, X, XI, respectively) (Masaki et al., 1976, 1978; Castaño et al., 1990; Damude et al., 1990; Bouâlam et al., 1992; Schmiedgen et al., 1994; Couce et al., 1996; Huber et al., 1997; de Castro et al., 2001; Sousa-Pedrares et al., 2003; Ismaylova et al., 2012), thus underlining their already intriguing ligand features in mononuclear group 14 element coordination chemistry. In addition, HPyS has been shown to enter the tin coordination sphere as a charge-neutral ligand via S-donor action (Valle et al., 1987; Wu et al., 2000). Therefore, this article will deal with the molecular structures of some analogous Si and Ge compounds and some hitherto not crystallographically characterized Sn compounds of PyS-.

PyS-bridged heteronuclear group 14 complexes.

Different coordination patterns of Sn(PyS)-complexes.
Results and discussion
With focus on the utilization of (PyS) and chlorine-substituted derivatives of Sn, Ge, and Si in further studies (i.e., syntheses of heteronuclear complexes of group 14 and, e.g., group 10 elements), we aimed at the syntheses and characterization of the compounds E(PyS)4 and Cl2E(PyS)2 (E=Sn, Ge, Si). Although the tin compounds Sn(PyS)4 and Cl2Sn(PyS)2 have already been described in the literature (Masaki et al., 1976; Damude et al., 1990), the crystal structure data of pure Sn(PyS)4 are lacking (only a structure of Sn(PyS)4·HPyS has been reported), and the crystal structure of Cl2Sn(PyS)2 in its triclinic modification (Masaki et al., 1978), which is related to our novel Ge and Si analogs, was of moderate quality only, and therefore, we included the synthesis of Sn(PyS)4 and structural characterization of Sn(PyS)4 and Cl2Sn(PyS)2 in this article as well. (Note: after submission of this manuscript, the synthesis and characterization of Cl2Si(PyS)2 were published by Baus et al., 2013; the experimental data reported therein are in accord with our data, even though the authors used a different synthesis route, i.e., starting from SiCl4, 2-mercaptopyridine, and triethylamine without a transsilylation step.)
The compounds Cl2E(PyS)2 (E=Ge, Si) were prepared along a transsilylation route (Scheme 4). This rather mild substitution approach gave rise to the desired products in good yields without the risk of threefold or fourfold substitution of the Ge- or Si-bound chlorine atoms. We found all compounds of the series Cl2E(PyS)2 (E=Sn, Ge, Si) to crystallize in the triclinic system with similar unit cell parameters (Table 1). As a prerequisite, the crystal structures were founded on molecules of similar complex architecture (Figure 1). Selected bond lengths and angles are listed in Table 2 for comparison. The E-N and E-S bond lengths were in accord with the increasing atomic radii of the central atoms of the complexes. Noteworthy, an increase in thiophilicity was observed from the Si toward the Ge compound. That is, whereas the Ge-N bonds (2.09 and 2.10 Å) were ca. 0.15 Å longer than the Si-N bonds (1.95 Å), less bond length increase was found for the Ge-S bonds (2.32 Å), which were ca. 0.05 Å longer than the Si-S bonds (2.27 Å). In the transition from the Ge to the Sn compound, both the E-N and E-S bonds revealed similar trends (increase by ca. 0.16 and 0.15 Å, respectively; i.e., Sn-N 2.26 and 2.27 Å, Sn-S 2.47 and 2.48 Å). In the compounds Cl2E(PyS)2 (E=Sn, Ge, Si), the Cl-E-Cl angles were similar and slightly wider than 90° (ranging between 93° and 97°), which can be attributed to the low steric demand of the four-membered chelates. The cis-arranged nitrogen donor atoms furnished an N-E-N angle of close to 90° in all cases (ranging between 85° and 89°), whereas the S-E-S angles exhibited noticeable differences, i.e., pronounced deviation from linearity with an increasing atomic radius of E (163.3°, 159.4°, and 154.6° or 156.8° for Si, Ge, and Sn, respectively). In principle, some differences can be observed between the molecular structures of Cl2Sn(PyS)2 in its two modifications, which hinted at some flexibility of those molecules to respond to crystal packing effects. Probably most interesting was the N-C-S angle within each of the four-membered chelates, which was significantly smaller than 120° (ranging between 107° and 113°). Thus, this angle clearly indicated attraction between the central atom E and both donor atoms S and N of each chelate, rather than chelation as an exclusive result of spatial proximity of E and the N or S atom. Furthermore, the decrease in the N-C-S angle along the series E=Sn>Ge>Si was in support of this attractive interaction, as the E-S bond responded to the decreasing atomic radius of the central atom. This observation clearly distinguished the PyS- ligand from the related mt-, which furnished capped tetrahedral coordination spheres in its monosilane derivatives with rather long Si···S separations (>3 Å) (Wagler et al., 2010c).

Syntheses of (PyS)-complexes of silicon, germanium and tin.
Crystallographic parameters of data collection and structure refinement for the crystal structures of Cl2Si(PyS)2, Cl2Ge(PyS)2, Cl2Sn(PyS)2, and Si(PyS)4.
Cl2Si(PyS)2 | Cl2Ge(PyS)2 | Cl2Sn(PyS)2 | Si(PyS)4 | |
---|---|---|---|---|
Empirical formula | C10H8Cl2N2S2Si | C10H8Cl2GeN2S2 | C10H8Cl2N2S2Sn | C20H16N4S4Si |
Formula weight | 319.29 | 363.79 | 409.89 | 468.70 |
T (K) | 200 (2) | 150 (2) | 150 (2) | 150 (2) |
λ (Å) | 0.71073 | |||
Crystal system | Triclinic | Triclinic | Triclinic | Orthorhombic |
Space group | P-1 | P-1 | P-1 | Pbca |
Unit cell dimensions | ||||
a (Å) | 7.1433 (7) | 7.0950 (11) | 7.1186 (5) | 13.3590 (3) |
b (Å) | 8.0790 (7) | 8.0963 (11) | 8.3342 (6) | 16.3995 (4) |
c (Å) | 12.2237 (11) | 12.324 (3) | 12.4096 (9) | 19.2429 (7) |
α (°) | 82.464 (8) | 82.031 (15) | 80.694 (6) | 90 |
β (°) | 86.589 (8) | 86.678 (15) | 86.510 (6) | 90 |
γ (°) | 67.967 (7) | 68.585 (11) | 69.034 (5) | 90 |
V (Å3) | 648.23 (11) | 652.7 (2) | 678.44 (8) | 4215.8 (2) |
Z/Dc (g/cm3) | 2/1.636 | 2/1.851 | 2/2.007 | 8/1.477 |
μ (mm-1) | 0.891 | 3.052 | 2.562 | 0.523 |
F (000) | 324 | 360 | 396 | 1936 |
Crystal size (mm) | 0.30×0.30×0.05 | 0.35×0.20×0.08 | 0.40×0.20×0.10 | 0.35×0.18×0.10 |
θ range for data collection | 2.7–32.0 | 2.7–32.0 | 2.7–32.0 | 2.5–30.0 |
Reflections collected | 12 058 | 21 575 | 16 220 | 70 279 |
Independent reflections/Rint | 4473/0.0307 | 4533/0.0314 | 4709/0.0317 | 6143/0.0388 |
Completeness to θmax | 99.4% | 100% | 100% | 99.9% |
Refinement | Full-matrix least-squares on F2 | |||
Data/restraints/parameters | 4473/0/154 | 4533/0/154 | 4709/0/154 | 6143/1/262 |
Goodness of fit on F2 | 1.074 | 1.056 | 1.056 | 1.072 |
R1/wR2 [I>2σ(I)] | 0.0333/0.0693 | 0.0191/0.0415 | 0.0155/0.0363 | 0.0271/0.0661 |
R1/wR2 (all data) | 0.0534/0.0786 | 0.0232/0.0424 | 0.0183/0.0371 | 0.0322/0.0683 |
Largest diff. peak and hole, eÅ-3 | 0.556/-0.333 | 0.435/-0.258 | 0.496/-0.509 | 0.420/-0.278 |
![Figure 1 Molecular structures of Cl2E(PyS)2 (E=Si, Ge, Sn) in their crystal structures.ORTEP/POV-RAY diagrams (Farrugia, 1997; Persistence of Vision Pty. Ltd. Persistence of Vision Raytracer Version 3.6 [POV-RAY 3.6], 2004) with thermal displacement ellipsoids are shown at the 30% (Si compound) and 50% (Ge and Sn compound) probability levels. Selected atoms are labeled and hydrogen atoms are omitted for clarity.](/document/doi/10.1515/mgmc-2013-0041/asset/graphic/mgmc-2013-0041_fig1.jpg)
Molecular structures of Cl2E(PyS)2 (E=Si, Ge, Sn) in their crystal structures.
ORTEP/POV-RAY diagrams (Farrugia, 1997; Persistence of Vision Pty. Ltd. Persistence of Vision Raytracer Version 3.6 [POV-RAY 3.6], 2004) with thermal displacement ellipsoids are shown at the 30% (Si compound) and 50% (Ge and Sn compound) probability levels. Selected atoms are labeled and hydrogen atoms are omitted for clarity.
Selected bond lengths (Å) and angles (°) in the crystal structures of compounds Cl2E(PyS)2 (E=Si, Ge, Sn).
Cl2Si(PyS)2a | Cl2Ge(PyS)2a | Cl2Sn(PyS)2a | Cl2Sn(PyS)2b | |
---|---|---|---|---|
E(1)-N(1) | 1.951 (2) | 2.092 (1) | 2.265 (1) | 2.259 (2) |
E(1)-N(2) | 1.953 (2) | 2.100 (1) | 2.263 (1) | |
E(1)-Cl(1) | 2.1675 (7) | 2.2575 (6) | 2.3969 (4) | 2.3892 (8) |
E(1)-Cl(2) | 2.1890( 7) | 2.2707 (5) | 2.4080 (4) | |
E(1)-S(1) | 2.2661 (6) | 2.3237 (5) | 2.4739 (3) | 2.4779 (8) |
E(1)-S(2) | 2.2690 (6) | 2.3217 (4) | 2.4770 (4) | |
N(1)-E(1)-N(2) | 88.79 (6) | 88.15 (4) | 88.23 (4) | 85.7 (1) |
Cl(1)-E(1)-Cl(2) | 93.63 (3) | 94.54 (2) | 94.59 (1) | 96.36 (4) |
S(1)-E(1)-S(2) | 163.25 (3) | 159.37 (1) | 154.55 (1) | 156.76 (4) |
N(1)-C(1)-S(1) | 107.8 (1) | 109.93 (8) | 112.81 (9) | 112.7 (2) |
N(2)-C(6)-S(2) | 107.8 (1) | 109.86 (8) | 112.72 (9) | |
N(1)-E(1)-S(1) | 72.78 (4) | 70.28 (3) | 66.26 (3) | 65.85 (7) |
N(2)-E(1)-S(2) | 72.56 (5) | 70.26 (3) | 66.19 (3) | |
N(2)-E(1)-Cl(1) | 166.13 (5) | 165.10 (3) | 163.06 (3) | 159.13 (7) |
N(1)-E(1)-Cl(2) | 165.34 (5) | 163.74 (3) | 161.05 (3) |
aEntries in these columns relate to crystal structures determined in this work.
bEntries in this column refer to related parameters in the crystal structure of the monoclinic modification of Cl2Sn(PyS)2 (Ismaylova et al., 2012).
For the syntheses of compounds E(PyS)4 (E=Si, Ge, Sn), different synthetic approaches were used (Scheme 4). In case of compound Sn(PyS)4, an oxidative addition approach was successful (2,2′-dipyridyldisulfide is generally known to undergo oxidative addition with low-valent tin compounds; Masaki et al., 1976; Damude et al., 1990), which, in contrast to the original literature procedure (Damude et al., 1990), started from a tin(II) complex instead of tin powder, to furnish a product of good purity. A substitution route was chosen for the preparation of Si(PyS)4 because of the pronounced instability of Si(II) compounds. With respect to the related Ge complex, we have not been able to isolate Ge(PyS)4. So far, both routes (oxidative addition and substitution) have led to mixtures of products. Thus, only the molecular structures of E(PyS)4 (E=Si, Sn) will be discussed in the following (Tables 1, 3, and 4, Figure 2). Furthermore, from the filtrate of the synthesis of Sn(PyS)2, some crystals of a new modification of Sn(PyS)4·HPyS (which actually is a decomposition product) were obtained, and thus, this structure will be included in the comparison (Figure 3).

Molecular structures of E(PyS)4 (E=Si, Sn) in their crystal structures.
ORTEP/POV-RAY diagrams (Farrugia, 1997; POV-RAY 3.6, 2004) with thermal displacement ellipsoids are shown at the 50% probability level. Selected atoms are labeled and hydrogen atoms are omitted for clarity. (Note: The labels of S3,C11,N3 and S4,C16,N4 in compound Sn(PyS)4 do not correspond to the labels in the CIF but have been assigned for consistence with the corresponding bonds and angles in Table 3.)

Molecular structures of Sn(PyS)4·HPyS in the crystal structure.
ORTEP/POV-RAY diagrams (Farrugia, 1997; POV-RAY 3.6, 2004) with thermal displacement ellipsoids are shown at the 50% probability level. Selected atoms are labeled and C-bound hydrogen atoms are omitted for clarity.
Selected bond lengths and other interatomic separations (Å) and angles (°) in the crystal structures of compounds E(PyS)4 (E=Si, Sn) and related parameters from the Sn(PyS)4molecules of two different crystal structures of Sn(PyS)4·HPyS.
Si(PyS)4a | Sn(PyS)4a | Sn(PyS)4·HPySa | Sn(PyS)4·HPySb | |
---|---|---|---|---|
E(1)-N(1) | 1.950 (1) | 2.258 (1) | 2.325 (3) | 2.332 (5) |
E(1)-N(2) | 1.943 (1) | 2.293 (1) | 2.304 (3) | 2.324 (5) |
E(1)···N(3) | 3.524 (1) | 3.271 (1) | 3.017 (3) | 4.005 (5) |
E(1)···N(4) | 3.972 (1) | 4.163 (1) | 4.088 (3) | 4.626 (5) |
E(1)-S(1) | 2.3357 (4) | 2.5512 (4) | 2.543 (1) | 2.542 (2) |
E(1)-S(2) | 2.2901 (4) | 2.4975 (4) | 2.493 (1) | 2.483 (2) |
E(1)-S(3) | 2.2369 (4) | 2.4900 (4) | 2.469 (1) | 2.475 (2) |
E(1)-S(4) | 2.2547 (4) | 2.4684 (4) | 2.458 (1) | 2.463 (2) |
S(1)-E(1)-S(2) | 157.95 (2) | 148.10 (1) | 146.36 (4) | 148.22 (6) |
N(1)-E(1)-S(4) | 168.84 (3) | 161.98 (3) | 162.29 (9) | 156.2 (1) |
N(2)-E(1)-S(3) | 172.52 (3) | 170.01 (3) | 161.33 (9) | 159.4 (1) |
N(1)-C(1)-S(1) | 108.69 (8) | 113.1 (1) | 113.2 (3) | 113.4 (5) |
N(2)-C(6)-S(2) | 108.50 (8) | 112.8 (1) | 112.7 (3) | 113.6 (5) |
N(3)-C(11)-S(3) | 119.19 (9) | 118.65 (11) | 116.7 (3) | 117.5 (5) |
N(4)-C(16)-S(4) | 116.99 (8) | 116.62 (11) | 116.3 (3) | 118.1 (5) |
aEntries in these columns relate to crystal structures determined in this work.
bEntries in this column refer to related parameters in the crystal structure of the previously reported modification of Sn(PyS)4·HPyS (Damude et al., 1990).
Crystallographic parameters of data collection and structure refinement for the crystal structures of Sn(PyS)4, Sn(PyS)4·HPyS, Sn(PyS)2, and [Sn(PyS)2(S)]2·THF.
Sn(PyS)4 | Sn(PyS)4·HpyS | Sn(PyS)2 | [Sn(PyS)2(S)]2·THF | |
---|---|---|---|---|
Empirical formula | C20H16N4S4Sn | C25H21N5S5Sn | C40H32N8S8Sn4 | C24H24N4OS6Sn2 |
Formula weight | 559.30 | 670.46 | 1356.14 | 814.12 |
T (K) | 200 (2) | 200 (2) | 150 (2) | 200 (2) |
λ (Å) | 0.71073 | |||
Crystal system | Orthorhombic | Monoclinic | Monoclinic | Monoclinic |
Space group | Pbca | C2/c | P21/n | C2/c |
Unit cell dimensions | ||||
a (Å) | 14.1117 (5) | 23.7709 (7) | 8.0209 (3) | 17.1867 (11) |
b (Å) | 16.6798 (7) | 15.9729 (6) | 10.4294 (4) | 8.3003 (4) |
c (Å) | 18.7432 (6) | 16.3625 (5) | 27.6975 (13) | 23.9045 (16) |
α (°) | 90 | 90 | 90 | 90 |
β (°) | 90 | 116.147 (2) | 91.429 (4) | 116.355 (5) |
γ (°) | 90 | 90 | 90 | 90 |
V (Å3) | 4411.8 (3) | 5576.9 (3) | 2316.26 (16) | 3055.6 (3) |
Z/Dc (g/cm3) | 8/1.684 | 8/1.597 | 2/1.944 | 4/1.770 |
μ (mm-1) | 1.552 | 1.315 | 2.533 | 2.070 |
F (000) | 2224 | 2688 | 1312 | 1600 |
Crystal size (mm) | 0.40×0.30×0.25 | 0.20×0.10×0.10 | 0.30×0.10×0.04 | 0.30×0.25×0.09 |
θ range for data collection | 2.4–28.0 | 2.6–27.0 | 2.4–28.0 | 2.5–28.0 |
Reflections collected | 43 737 | 6093 | 16 734 | 16 653 |
Independent reflections/Rint | 5313/0.0304 | 6093/0.0000a | 5581/0.0255b | 3680/0.0406 |
Completeness to θmax | 99.9% | 99.9% | 99.8% | 99.7% |
Refinement | Full-matrix least-squares on F2 | |||
Data/restraints/parameters | 5313/0/262 | 6093/0/329 | 5581/0/272 | 3680/33/188 |
Goodness of fit on F2 | 1.055 | 1.122 | 1.115 | 1.052 |
R1/wR2 [I>2σ(I)] | 0.0172/0.0412 | 0.0480/0.1033 | 0.0269/0.0650 | 0.0224/0.0496 |
R1/wR2 (all data) | 0.0207/0.0423 | 0.0656/0.1103 | 0.0300/0.0671 | 0.0307/0.0521 |
Largest diff. peak and hole, eÅ-3 | 0.301/-0.265 | 1.004/-0.815 | 0.462/-0.395 | 0.405/-0.320 |
aAbsorption correction of this data set was performed with XABS2 as implemented in WinGX (Farrugia, 1999), which produced a merged data set. Number of independent reflections/Rint before XABS2 was 27 496/0.1154.
bThe data set was collected from a slightly twinned crystal. Upon data integration of the predominant population and initial refinement, the data set was detwinned using ROTAX as implemented in WinGX (Farrugia, 1999). The twin law (1 0 0)(0 -1 0)(-0.172 0 -1) corresponds to a 180° rotation about the [1 0 0] direct lattice direction. A merged HKLF5 file was created in WinGX for final refinement. The twin population parameter BASF refined to 0.0326 (3). The Rint reported was taken from the initial data set (with 16 751 collected reflections) prior to detwinning.
In general, the pronounced thiophilicity of tin over silicon was again reflected in the molecular structures of E(PyS)4 (E=Si, Sn), with the E-N bonds (1.94 and 1.95 Å for E=Si and 2.26 and 2.29 Å for E=Sn) being ca. 0.31–0.35 Å longer in the Sn complex, whereas E-S bond length increased by only 0.20–0.22 Å (2.29 and 2.34 Å for E=Si and 2.50 and 2.55 Å for E=Sn) and by 0.23 Å (2.24 and 2.25 Å for E=Si and 2.47 and 2.49 Å for E=Sn) for the chelate and nonchelate S atoms, respectively. Once again, the noticeably small N-C-S angles within the chelates (ca. 109° for E=Si, ca. 113° for E=Sn), which were smaller for the Si compound, reflected the attraction between the central atom E and both donor atoms of the chelates. In case of compounds E(PyS)4 (E=Si, Sn), the nonchelate (thus S-bound) (PyS) groups exhibited significantly larger N-C-S angles, which ranged between 116° and 120°. Interestingly, in the three different crystal structures of Sn(PyS)4 and Sn(PyS)4·HPyS, the molecule of Sn(PyS)4 exhibited different coordination modes. Even though the fundamental motif was a distorted octahedral coordination sphere about tin, with the monodentate ligands’ S atoms cis and the chelate S atoms trans, the Sn(PyS)4 molecules in Sn(PyS)4 and in the herein reported structure of Sn(PyS)4·HPyS showed [6+1] coordination of the tin atom, furnished by capping of one octahedral face by one of the nonchelate N atoms (N3). The capping modes were different from each other, as in Sn(PyS)4, a NS2 face was capped (N1,S2,S3), whereas in the related molecule in the structure of Sn(PyS)4·HPyS, an S3 face was capped (S2,S3,S4).
So far, Cl2Si(PyS)2, Cl2Ge(PyS)2, and Si(PyS)4 were the first crystallographically characterized representatives of Si(Cl2N2S2), Ge(Cl2S2N2), and Si(N2S4) coordination spheres, respectively (ConQuest 1.15, 2012). (The crystal structures of Cl2Si(PyS)2 and Cl2Ge(PyS)2 have been part of a presentation on the GTL2013 conference; Wächtler et al., 2013.) Whereas Si(PyS)4 exhibited sufficient solubility in CDCl3, which allowed for its 29Si NMR characterization in solution (the 29Si shift of -165 ppm clearly indicated that hexacoordination of the Si atom was retained in the solution, although the detection of a sole set of broad signals in the 1H and 13C NMR spectra indicated rapid exchange between the chelating and the nonchelating PyS moieties), the solubility of Cl2Si(PyS)2 was too poor for the acquisition of decent 29Si NMR spectroscopic data from a solution. Thus, 29Si solid-state NMR spectroscopy was employed. To our surprise, the 29Si CP/MAS NMR spectrum of Cl2Si(PyS)2 (shown in Figure 4) exhibited a broad signal that was patterned with two peaks and a shoulder. As the crystal structure of this complex would give rise to the expectation of a singlet signal, the pattern observed in the spectrum was interpreted as the result of dipolar coupling with the chlorine atoms, since similar effects have already been observed and successfully described for 31P solid-state NMR spectra (Thomas et al., 2001). Indeed, simulation of the signal with model ‘THREE’ in CSOLIDS (Eichele, 2013) produced an excellent explanation for this broad and patterned 29Si resonance. The following parameters were used to describe the couplings with the 29Si nucleus: dipolar coupling with 35Cl, which is representative for Cl, as it was shown to be similar to 37Cl (Brendler et al., 2012); D=-231.058 and -224.585 Hz for the two independent crystallographic chlorine sites (values derived from the crystallographically determined Si-Cl bond lengths); 1J(29Si-35Cl)=30 Hz as an initial guess (Brendler et al., 2012; Wagler et al., 2013); quadrupolar couplings Cq=29 and 34 MHz for the two independent chlorine sites (Cq was varied to achieve good fit of the simulation); Cl-Si-Cl angle 93.5° (derived from the X-ray data); and a line broadening GB=75 Hz.

29Si CP/MAS NMR spectrum of Cl2Si(PyS)2.
Signal traces are as follows (from top): simulated signal with smoothing (red), experimental spectrum (blue), and simulated signal as stick approximation without smoothing (green).
As the tin(II) compound Sn(PyS)2 (Ichikawa and Mukaiyama, 1985) provided easy access to Sn(PyS)4 via oxidative addition of a disulfide, the synthesis potential of Sn(PyS)2 was further probed in the reaction with elemental sulfur. As expected, a tin(IV) compound of the gross composition SnS(PyS)2 formed. Revealed by X-ray crystallography (Figure 5), this compound is a cyclic dimer. Due to its location around a crystallographic center of inversion, the Sn2S2 four-membered ring was constrained to perfect planarity. Furthermore, the anisotropic displacement parameters of the atoms of this cycle hinted at only limited displacement out of this perfectly planar arrangement by effects such as thermal motion or dynamic disorder, and the two independent Sn-S bonds within the Sn2S2 four-membered ring exhibited similar lengths (2.45 and 2.46 Å). This Sn2S2 arrangement with hexacoordinate Sn was similar to the related motif in other compounds (e.g., Scheme 5, XII and XIII), which have previously been described in the literature (Jurkschat et al., 1992; Nayek et al., 2009; Rotar et al., 2009).
![Figure 5 Molecular structure of [SnS(PyS)2]2 in the crystal structure of [SnS(PyS)2]2·THF.ORTEP/POV-RAY diagrams (Farrugia, 1997; POV-RAY 3.6, 2004) with thermal displacement ellipsoids are shown at the 50% probability level. Selected atoms are labeled and hydrogen atoms and solvent molecule are omitted for clarity. The asymmetric unit comprises one half of a centrosymmetric cyclo-dimer. Symmetry-related atomic sites are indicated by * (for inversion 1-x, -y, -z). Selected bond lengths (Å) and angles (°): Sn1-S1 2.5483 (7), Sn1-S2 2.5394 (6), Sn1-S3 2.4456 (6), Sn1-S3* 2.4615 (7), Sn1-N1 2.293 (2), Sn1-N2 2.266 (2), Sn1-S3-Sn1* 86.43 (2), S3-Sn-S3* 93.57 (2), S1-Sn1-S2 145.20 (2), N1-Sn1-N2 86.05 (7), N1-C1-S1 113.2 (2), N2-C6-S2 113.0 (2).](/document/doi/10.1515/mgmc-2013-0041/asset/graphic/mgmc-2013-0041_fig5.jpg)
Molecular structure of [SnS(PyS)2]2 in the crystal structure of [SnS(PyS)2]2·THF.
ORTEP/POV-RAY diagrams (Farrugia, 1997; POV-RAY 3.6, 2004) with thermal displacement ellipsoids are shown at the 50% probability level. Selected atoms are labeled and hydrogen atoms and solvent molecule are omitted for clarity. The asymmetric unit comprises one half of a centrosymmetric cyclo-dimer. Symmetry-related atomic sites are indicated by * (for inversion 1-x, -y, -z). Selected bond lengths (Å) and angles (°): Sn1-S1 2.5483 (7), Sn1-S2 2.5394 (6), Sn1-S3 2.4456 (6), Sn1-S3* 2.4615 (7), Sn1-N1 2.293 (2), Sn1-N2 2.266 (2), Sn1-S3-Sn1* 86.43 (2), S3-Sn-S3* 93.57 (2), S1-Sn1-S2 145.20 (2), N1-Sn1-N2 86.05 (7), N1-C1-S1 113.2 (2), N2-C6-S2 113.0 (2).

Selected examples of hexacoordinate tin compounds with Sn2S2 motif.
The S3-Sn1-S3* angle (93.6°) was similar to the Cl-Sn-Cl angle in Cl2Sn(PyS)2 (94.6° and 96.4°), and the same was true for the N-Sn-N angle (86.1° in [SnS(PyS)2]2, 88.2° and 85.7° in Cl2Sn(PyS)2). Interestingly, the angle S1-Sn1-S2 (145.2°) deviated far more from linearity than in the compound Cl2Sn(PyS)2 (154.6° and 156.8°), which can be attributed to the longer Sn1-S1 and Sn1-S2 bonds in (SnS(PyS)2)2 (2.55 and 2.54 Å). Apparently, the decrease in Lewis acidity (caused by substitution of the Sn-bound Cl atoms for S atoms) caused predominant lengthening of the Sn-S bonds to the chelating ligands. At first glance, this might be unexpected in the context of the previously mentioned pronounced thiophilicity of Sn (pronounced attraction of S over N donors), but the canonical form of PyS- as a 1,2-dihydropyridine-2-thione anion offered an explanation with the formally dative (thus more flexible) Sn-S bond within the chelate. This is still unexpected because in compounds such as VIII and X (Scheme 3), the N atoms appeared to establish the formally dative (weaker, and thus more flexible) bond.
Finally, the solid-state structure of the tin(II) compound Sn(PyS)2, which had been reported in the literature (Ichikawa and Mukaiyama, 1985) and was used for the synthesis of Sn(PyS)4, has not been reported yet. Thus, we collected the single-crystal X-ray diffraction data set of this compound, and structure solution and refinement revealed an interesting arrangement of the molecules in the solid state (Figure 6). In principle, the crystal structure comprised two independent Sn(PyS)2 moieties. Interestingly, one of them was part of a centrosymmetric cyclo-dimer, whereas the other moiety was part of a polymeric chain (which was aligned along the monoclinic b axis around a 21 screw axis). Both tin atoms were [3+2] coordinated, but the one in the chain comprised a [SN2+S2] coordination, whereas in the cyclo-dimer, the tin coordination sphere was [S2N+SN]. Their same coordination number was clearly reflected by their similar 119Sn NMR shifts in the solid state (-353 and -368 ppm). The 119Sn chemical shift tensors, however, exhibited notable differences (Figure 7). In the CDCl3 solution, the additional lone pair donation was absent (or at least significantly less pronounced), as reflected by a noticeable downfield shift of the 119Sn resonance (-206 ppm), and same coordination mode (or at least rapid exchange) caused the presence of only one set of 1H and 13C NMR signals for the (PyS) ligands. This structure once more underlined the coordinative flexibility of the pyridine-2-thiolato ligand, which may thus adapt not only to the electronic features of a central atom of a complex but also to crystal packing environments by operation via pronounced S- or N-donor activity.

Molecular structure of Sn(PyS)2 in the crystal structure.
ORTEP/POV-RAY diagrams (Farrugia, 1997; POV-RAY 3.6, 2004) with thermal displacement ellipsoids are shown at the 50% probability level. Selected atoms are labeled and hydrogen atoms are omitted for clarity. The asymmetric unit comprises two independent Sn(PyS)2 moieties. Symmetry-related atomic sites are indicated by * (for the shift along a 21 axis 1.5-x, y+0.5, 0.5-z) or ** (for inversion 2-x, 2-y, -z). Selected bond lengths (Å): Sn1-S1 2.6389 (5), Sn1-S2 3.0351 (6), Sn1-S2* 3.2370 (6), Sn1-N1 2.332 (2), Sn1-N2 2.268 (2), Sn2-S3 2.5709 (5), Sn2-S4 3.2182 (5), Sn2-S4** 2.6627 (5), Sn2-N3 2.672 (2), Sn2-N4 2.328 (2).

119Sn MAS NMR spectrum of Sn(PyS)2.
Signal traces are as follows (from top): simulated signal 1 (purple), simulated signal 2 (green), simulated spectrum with signals 1 and 2 (red), and experimental spectrum (blue). The characteristics of the chemical shift anisotropy tensors are as follows: for signal 1, δiso -368.2 ppm, δ11 -57.1 ppm, δ22 -112.2 ppm, δ33 -935.3 ppm, Ω 878.2 ppm, κ 0.87; for signal 2, δiso -353.0 ppm, δ11 110.0 ppm, δ22 -315.1 ppm, δ33 -853.9 ppm, Ω 963.8 ppm, κ 0.12.
Conclusions
In this article, we have presented the first comparison of structurally related Si, Ge, and Sn compounds with respect to the coordination chemistry of pyridine-2-thiolate, which was enabled by the first crystallographically characterized Si and Ge complexes of this interesting ambidentate ligand. Furthermore, crystallographic characterization of some literature known pyridine-2-thiolato tin complexes underlined the coordinative flexibility and versatility of this ligand in group 14 coordination chemistry. The herein reported Si, Ge, and Sn compounds are currently under investigation as starting materials for the syntheses of heteronuclear complexes with group 14 elements and transition metals.
Experimental
General considerations and analyses
Most of the compounds are sensitive toward air and moisture. Thus, all reactions were routinely carried out using standard Schlenk technique. The solvents and liquid starting materials used were dried as follows: THF, diethyl ether, and triethylamine: distillation from sodium-benzophenone; toluene and pentane: distillation from sodium; and chloroform (CDCl3 and amylene stabilized CHCl3) and deuterated dimethyl sulfoxide: molecular sieves 3 Å. SnCl2(dioxane) was prepared according to a previously reported procedure (Morrison and Haendler, 1967), and all other chemicals were commercially available and used as received.
C/H/N mircoanalyses were recorded on a ‘Vario Micro Cube’ analyzer (Elementar, Hanau, Germany). Melting points were determined in sealed glass capillaries on a Boëtius-type heating microscope (‘Polytherm A’, Wagner & Munz, München, Germany). NMR spectra were recorded on a DPX 400 spectrometer (Bruker Biospin, Germany, Rheinstetten) for 1H (400.1 MHz), 13C (100.6 MHz), and 29Si (79.5 MHz) or on an Avance 500 spectrometer (Bruker Biospin, Germany, Rheinstetten) for 119Sn (186.5 MHz) as CDCl3 solutions (unless otherwise stated). The spectra are referenced against internal SiMe4 (1H, 13C, and 29Si) or external SnMe4 (119Sn). Chemical shifts are given in ppm. 119Sn MAS NMR (149.2 MHz) and 29Si CP/MAS NMR (79.5 MHz) spectra were recorded on an Avance 400 WB spectrometer (Bruker Biospin, Germany, Rheinstetten) using 4-mm zirconia (ZrO2) spinners (119Sn) and 7-mm zirconia spinners (29Si). Solid-state NMR spectra are referenced against SnO2 (-603 ppm) for 119Sn and against octakis(trimethylsilyl)silsesquioxane (Q8M8; -109 ppm for most shielded Q-group) for 29Si. DMFIT (Massiot et al., 2002) and HBMAS (Fenzke, 1989) were used to calculate the principal components of the shielding tensor from spinning sideband spectra. The residual dipolar couplings between silicon and chlorine were simulated using the WSOLIDS NMR simulation package (Eichele, 2013). Crystals suitable for X-ray diffraction were selected under an inert oil and mounted on a glass capilliary with a thin film of silicon grease. The data sets were collected on an IPDS-2(T) diffractometer (Stoe, Darmstadt, Germany). Structures were solved by direct methods with ShelXS (Sheldrick, 1997b), and all nonhydrogen atoms were anisotropically refined in full-matrix least-squares cycles against |F2| (ShelXL) (Sheldrick, 1997a). Carbon-bound hydrogen atoms were placed in idealized positions and refined isotropically. In case of Sn(PyS4)·HPyS, the N-bound hydrogen atom was detected as a residual electron density peak. The structure of [Sn(PyS)2(S)]2·THF contains heavily disordered solvent molecules (disorder by symmetry around a twofold axis) and disorder of the solvent part of the asymmetric unit over two sites [refined with site occupancy factors 0.278 (5) and 0.222 (5)]. Crystallographic data for the structures reported in this article have been deposited with the Cambridge Crystallographic Data Center (CCDC) as supplementary material publication no. CCDC-956098 (Cl2Si(PyS)2), CCDC-956099 (Sn(PyS)4·HPyS), CCDC-956100 (Sn(PyS)2), CCDC-956101 (Cl2Ge(PyS)2), CCDC-956102 (Cl2Sn(PyS)2), CCDC-956103 (Sn(PyS)4), CCDC-956104 ([Sn(PyS)2(S)]2·THF), and CCDC-956105 (Si(PyS)4). Copies of the data can be obtained free of charge on application to CCDC, 12 Union Road, Cambridge, CB2 1EZ, UK (fax: + 44-1223-336033; e-mail: deposit@ccdc.cam.ac.uk).
Synthesis of Cl2Si(PyS)2
Chlorotrimethylsilane (1.00 g, 9.20 mmol) was slowly added to a cold (0°C) solution of 2-mercaptopyridine (1.00 g, 9.00 mmol) and triethylamine (1.53 g, 15.0 mmol) in THF (40 mL), whereupon a white precipitate of Et3NHCl formed. After stirring for 1 h at 0°C, the precipitate was filtered off and washed with THF (3 mL). From the filtrate, all volatiles were removed under reduced pressure and the residue was dissolved in THF (20 mL), followed by addition of SiCl4 (0.76 g, 4.50 mmol) via syringe. The mixture was stirred for a few seconds, after which the solution was allowed to stand for 5 days, thus leading to the formation of a white solid, which was filtered off, washed with THF, and dried under vacuum. Yield: 0.73 g (2.28 mmol, 51%). The complex has poor solubility in common organic solvents. Crystals suitable for X-ray crystallography were grown by slow diffusion of diethyl ether into a saturated THF solution. M.p.: 120°C (decomposition); 1H NMR (400.1 MHz, CDCl3) δ: 6.78 (m, 2H, N-CH-CH), 7.39–7.57 (m, 4H, N-CH-CH-CH and N-CH-CH-CH-CH), and 7.75 (m, 2H, N-CH). 29Si NMR (79.5 MHz, solid state) δiso: -180 (broad signal, patterned with two peaks). Elemental analyses: calculated (%) for C10H8N2S2Cl2Si (319.29 g/mol): C 37.61, H 2.52, N 8.77; found: C 37.3, H 2.5, N 8.5.
Synthesis of Si(PyS)4
A 2 M solution of NaN(SiMe3)2 in THF (3.16 g, ρ=0.916 g/mL, 6.90 mmol) was added via syringe to a solution of 2-mercaptopyridine (0.73 g, 6.57 mmol) in THF (10 mL), and the mixture was briefly stirred and then stored at room temperature. After 2 days, all volatiles were removed under vacuum, the orange solid residue was dissolved in THF (50 mL), and solid Cl2Si(PyS)2 (1.00 g, 3.13 mmol) was added. The suspension was stirred at room temperature for 20 h and then filtered. The filtrate was concentrated to a volume of 15 mL by vacuum condensation into a cold trap. Upon storage at room temperature, a crystalline solid formed, which was isolated by decantation, washed with THF (2 mL), and dried under vacuum. Yield: 1.03 g (2.19 mmol, 70%). The sample contained crystals suitable for X-ray analysis. M.p.: 190°C (decomposition); 1H NMR (400.1 MHz, CDCl3) δ: 6.83 (br, 4H, N-CH-CH), 7.36 (br, 4H), 7.44 (br, 4H), and 8.19 (br, 4H, N-CH). 13C{1H} NMR (100.6 MHz, CDCl3) δ: 118.9, 127.4, 138.1, 144.7 (br, pyridine C-H), and 163.4 (br, C-S). 29Si NMR (79.5 MHz, CDCl3) δ: -165.0. Elemental analyses: calculated (%) for C20H16N4S4Si (468.70 g/mol): C 51.25, H 3.44, N 11.95; found: C 50.9, H 3.7, N 11.6.
Synthesis of Cl2Ge(PyS)2
Chlorotrimethylsilane (1.00 g, 9.20 mmol) was slowly added to a cold (0°C) solution of 2-mercaptopyridine (1.00 g, 9.00 mmol) and triethylamine (1.53 g, 15.0 mmol) in THF (40 mL), whereupon a white precipitate of Et3NHCl formed. After stirring for 2 h at 0°C, the precipitate was filtered off and the volatiles of the filtrate were removed under vacuum. The residue was dissolved into chloroform (20 mL) and GeCl4 (1.00 g, 4.60 mmol) was added, whereupon the product started to precipitate immediately. After 1 h, the white solid was filtered off, washed with chloroform (5 mL), and dried under vacuum. Yield: 1.24 g (3.40 mmol, 76%). Crystals of this compound were grown by a slightly modified method, i.e., by addition of GeCl4 to a THF solution of the silylated 2-mercaptopyridine (cf. Synthesis of Cl2Si(PyS)2). Even though this alternative route was suitable for growing single crystals, the reaction was very slow and yield was very poor. M.p.: 238°C–244°C; 1H NMR (400.1 MHz, CDCl3) δ: 7.04 (m, 2H, N-CH-CH), 7.58–7.62 (m, 4H, N-CH-CH-CH and N-CH-CH-CH-CH), and 7.93 (m, 2H, N-CH). Elemental analyses: calculated (%) for C10H8N2S2Cl2Ge (363.79 g/mol): C 33.01, H 2.21, N 7.70; found: C 32.8, H 2.1, N 7.7.
Synthesis of Sn(PyS)2 (related to a literature procedure, Ichikawa and Mukaiyama, 1985)
Solid SnCl2(dioxane) (3.00 g, 10.8 mmol) was added to a stirred suspension of 2-mercaptopyridine (2.40 g, 21.6 mmol) in toluene (150 mL) and triethylamine (3.05 g, 30.0 mmol). The mixture was warmed slightly (ca. 50°C) under stirring for 5 h. After cooling to room temperature, the precipitate (Et3NHCl) was filtered off. Some solvent was removed from the filtrate by vacuum cold trap condensation and the resultant solution (volume of about 40 mL) was stored at -24°C for crystallization. After 5 days, the crystalline product was filtered off, washed with a mixture of toluene and pentane (2.5 mL each) and then with pentane (3 mL), and dried under vacuum. Yield: 2.62 g (7.73 mmol, 72%). The sample contained crystals suitable for X-ray diffraction. M.p.: 124°C–128°C; 1H NMR (400.1 MHz, CDCl3) δ: 6.93 (m, 2H, N-CH-CH), 7.17 (d, 3J(1H-1H)=8.1 Hz, 2H, N-CH-CH-CH-CH), 7.45 (m, 2H, N-CH-CH-CH), and 8.06 (d, 3J(1H-1H)=5.1 Hz, 2H, N-CH). 13C{1H} NMR (100.6 MHz, CDCl3) δ: 117.7, 128.1, 137.5, 145.4 (pyridine C-H), and 170.5 (C-S). 119Sn NMR (186.5 MHz, CDCl3) δ: -206.3; (149.2 MHz, solid state) δiso: -353.1, -368.3. Elemental analyses: calculated (%) for C10H8N2S2Sn (339.04 g/mol): C 35.43, H 2.38, N 8.26; found: C 35.3, H 2.4, N 8.2.
Synthesis of Sn(PyS)4
A suspension of Sn(PyS)2 (668 mg, 1.97 mmol) and 2,2′-dipyridyldisulfide (434 mg, 1.97 mmol) in toluene (5 mL) was stirred and heated to 110°C and more toluene (ca. 10 mL) was added until a clear solution was obtained. Upon cooling to room temperature, yellow crystals formed, which were filtered off, washed with toluene (4 mL), and dried under vacuum. Yield: 972 mg (1.74 mmol, 88%). The sample contained good-quality crystals for X-ray diffraction. The identity of the bulk sample was confirmed by comparison with 1H and 13C NMR spectroscopic data from the literature (Damude et al., 1990; Huber et al., 1997) 1H NMR (400.1 MHz, CDCl3) δ: 6.91 (m, 4H, N-CH-CH), 7.33 (d, 3J(1H-1H)=8.0 Hz, 4H, N-CH-CH-CH-CH), 7.43 (m, 4H, N-CH-CH-CH), and 8.04 (m, 4H, N-CH). 13C{1H} NMR (100.6 MHz, CDCl3) δ: 119.1, 125.8, 137.8, 145.7 (pyridine C-H), and 162.9 (C-S). 119Sn NMR (186.5 MHz, CDCl3) δ: -515.0. Elemental analyses: calculated (%) for C20H16N2S2Sn (559.30 g/mol): C 42.95, H 2.88, N 10.02; found: C 43.0, H 2.8, N 10.0.
Synthesis of [(PyS)2Sn(μ-S)]2
A suspension of Sn(PyS)2 (210 mg, 0.62 mmol) and sulfur (20 mg, 0.62 mmol) in THF (20 mL) was stirred at room temperature for 1 h and then at 60°C for 4 h. Thereafter, the mixture was allowed to cool to room temperature and the product was filtered off, washed with 1 mL of THF, and dried under vacuum. Yield: 188 mg (0.25 mmol, 82%). Crystals of the THF solvate were grown by layering Sn(PyS)2 with a THF solution of sulfur at room temperature. Within a few days, crystals suitable for X-ray diffraction analysis had formed. M.p.: no melting or decomposition observed up to 330°C; 1H NMR (400.1 MHz, (CD3)2SO) δ: 7.26 (t, 3J(1H-1H)=7.0 Hz, 4H, N-CH-CH), 7.54 (d, 3J(1H-1H)=8.1 Hz, 4H, N-CH-CH-CH-CH), 7.87 (m, 4H, N-CH-CH-CH), and 8.06 (d, 3J(1H-1H)=5.4 Hz, 4H, N-CH, satellites: 3J(1H-119/117Sn)=34 Hz, average for both isotopes, individual satellites not resolved). 13C{1H} NMR (100.6 MHz, (CD3)2SO) δ: 119.5, 123.6, 141.3, and 143.1 (pyridine C-H), 165.1 (C-S). 119Sn NMR (186.5 MHz, (CD3)2SO) δ: -498.0. Elemental analyses: calculated (%) for C20H16N4S6Sn2 (742.18 g/mol): C 32.37, H 2.17, N 7.55; found: C 32.5, H 2.3, N 7.3.
References
Autschbach, J.; Sutter, K.; Truflandier, L. A.; Brendler, E.; Wagler, J. Atomic Contributions from spin-orbit coupling to 29Si NMR chemical shifts in metallasilatrane complexes. Chem. Eur. J.2012, 18, 12803–12813.Search in Google Scholar
Baus, J. A.; Burschka, C.; Bertermann, R.; Fonseca Guerra, C.; Bickelhaupt, F. M.; Tacke, R. Neutral six-coordinate and cationic five-coordinate silicon(IV) complexes with two bidentate monoanionic N,S-pyridine-2-thiolato(-) ligands. Inorg. Chem.2013, 52, 10664–10676.Search in Google Scholar
Berrenguer, J. R.; Lalinde, E.; Martín, A.; Moreno, M. T.; Ruiz, S.; Sánchez, S.; Shahsavari, H. R. Solvent-induced lone pair activity tuning and vapoluminescence in a Pt2Pb cluster. Chem. Commun.2013, 49, 5067–5069.Search in Google Scholar
Bouâlam, M.; Meunier-Piret, J.; Biesemans, M.; Willem, R.; Gielen, M. Organotin(IV) compounds of 2-thiopyridine. Crystal and molecular structure of dicyclohexyltin(IV) bis(2-pyridylthiolate). Inorg. Chim. Acta1992, 198–200, 249–255.10.1016/S0020-1693(00)92367-3Search in Google Scholar
Brendler, E.; Wächtler, E.; Heine, T.; Zhechkov, L.; Langer, T.; Pöttgen, R.; Hill, A. F.; Wagler, J. Stannylene or metallastanna(IV)-ocane – a matter of formalism. Angew. Chem. Int. Ed.2011, 50, 4696–4700.Search 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.Search in Google Scholar
Castaño, M. V.; Macías, A.; Castiñeiras, A.; Sánchez González, A.; García Martínez, E.; Casas, J. S.; Sordo, J. Comparative structural study of dimethyl(pyridine-2-t hiolato)thallium(III) and dimethylbis(pyridine-2-thiolato)tin(IV). J. Chem. Soc. Dalton Trans.1990, 1001–1005.10.1039/DT9900001001Search in Google Scholar
ConQuest 1.15. Search of the Cambridge Crystal Structure Database (CSD). 2012. The Cambridge Crystallographic Data Centre, Cambridge, UK.Search in Google Scholar
Couce, M. D.; Cherchi, V.; Faraglia, G.; Russo, U.; Sindellari, L.; Valle, G.; Zancan, N. Synthesis and characterisation of organotin complexes with 2-mercaptopyridine. Appl. Organomet. Chem.1996, 10, 35–45.Search in Google Scholar
Damude, L. C.; Dean, P. A. W.; Manivannan, V.; Srivastava, R. S.; Vittal, J. J. Synthesis and multinuclear magnetic resonance study of some tin(IV) complexes of pyridine-2-thiolate and related ligands, and the X-ray structural analysis of Sn(C5H4NS)4·C5H5NS. Can. J. Chem.1990, 68, 1323–1331.Search in Google Scholar
de Castro, V. D.; Filgueiras, C. A. L.; de Lima, G. M.; Gambardella, M. T. P.; Speziali, N. L. Reactivity of [SnPh3(η1-2-PyS)] towards Pt and Ni complexes. Main Group Met. Chem.2001, 24, 761–764.Search in Google Scholar
Eichele, K. CSOLIDS; WSOLIDS, Solid-State NMR Spectrum Simulation; Universität Tübingen: Tübingen, Germany, 2013.Search in Google Scholar
Farrugia, L. J. ORTEP-3 for Windows – a version of ORTEP-III with a graphical user interphase (GUI). J. Appl. Crystallogr.1997, 30, 565.Search in Google Scholar
Farrugia, L. J. WINGX (version 1.64.05) An integrated system of windows programs fort he solution, refinement and analysis of single crystal X-ray diffraction data. J. Appl. Crystallogr.1999, 32, 837–838.Search in Google Scholar
Fenzke, D. HBMAS; Universität Leipzig: Leipzig, Germany, 1989.Search in Google Scholar
Huber, F.; Schmiedgen, R.; Schürmann, M.; Barbier, R.; Ruisi, G.; Silvestri, A. Mono-organotin(IV) and tin(IV) derivatives of 2-mercaptopyridine and 2-mercaptopyrimidine: X-ray structures of methyl-tris(2-pyridinethiolato)tin(IV) and phenyl-tris(2-pyridinethiolato)tin(IV)·1.5CHCl3. Appl. Organomet. Chem.1997, 11, 869–888.Search in Google Scholar
Ichikawa, J.; Mukaiyama, T. A facile synthesis of α,β-dihydroxyketones via tin(IV) enediolates by utilizing the reducing power of bis(2-pyridinethiolato)tin(II). Chem. Lett.1985, 1009–1012.10.1246/cl.1985.1009Search in Google Scholar
Ismaylova, S. R.; Matsulevich, Z. V.; Borisova, G. N.; Borisov, A. V.; Khrustalev, V. N. Dichlorobis(pyridine-2-thiolato-κ2N,S)-tin(IV): a new polymorph. Acta Crystallogr.2012, E68, m875–m876.Search in Google Scholar
Jurkschat, K.; van Dreumel, S.; Dyson, G.; Dakternieks, D.; Bastow, T. J.; Smith, M. E.; Dräger, M. Ring size control in diorganotin sulphides by intramolecular Sn-N coordination. Polyhedron1992, 11, 2747–2755.Search in Google Scholar
Martincová, J.; Dostál, L.; Herres-Pawlis, S.; Růžička, A.; Jambor, R. Intramolecularly coordinated [{2,6-(Me2NCH2)2C6H3}SnII]+: a strong σ donor for PtII. Chem. Eur. J.2011, 17, 7423–7427.Search in Google Scholar
Martincová, J.; Dostál, L.; Růžička, A.; Herres-Pawlis, S.; Jambor, R. Stabilization of an intramolecularly coordinated stannylidenium yation. Z. Anorg. Allg. Chem.2012, 638, 1672–1675.Search in Google Scholar
Masaki, M.; Matsunami, S. Synthesis of yihalogenobis(2-pyridinethiolato)tin(IV) by oxidative addition of disulfide to tin(II) halides – consideration on the coordination structure. Bull. Chem. Soc. Jpn.1976, 49, 3274–3279.Search in Google Scholar
Masaki, M.; Matsunami, S.; Udea, H. The crystal and molecular structure of dichlorobis(2-pyridinethiolato)tin(IV). Bull. Chem. Soc. Jpn.1978, 51, 3298–3301.Search in Google Scholar
Massiot, D.; Fayon, F.; Capron, M.; King, I.; Le Calvé, S.; Alonso, B.; Durand, J.-O.; Bujoli, B.; Gan, Z.; Hoatson, G. Modelling one- and two-dimensional solid stata NMR spectra. Magn. Reson. Chem.2002, 40, 70–76.Search in Google Scholar
Morrison, J. S.; Haendler, H. M. Some reactions of tin(II) chloride in nonaqueous solution. J. Inorg. Nucl. Chem.1967, 29, 393–400.Search in Google Scholar
Nayek, H. P.; Lin, Z.; Dehnen, S. Solvent-modified SiS2-type SnS2: synthesis, crystal structures and properties of [SnS2·en] and [enH]4[Sn2S6]·en. Z. Anorg. Allg. Chem.2009, 635, 1737–1740.Search in Google Scholar
Persistence of Vision Pty. Ltd. Persistence of Vision Raytracer (Version 3.6) [Computer software]. Retrieved from http://www.povray.org/download/, 2004.Search in Google Scholar
Rotar, A.; Varga, R. A.; Jurkschat, K.; Silvestru, C. Diorganotin(IV) compounds containing 2-(Et2NCH2)C6H4 moieties: configurational stability in solution and solid state structures. J. Organomet. Chem.2009, 694, 1385–1392.Search in Google Scholar
Schmiedgen, R.; Huber, F.; Preut, H.; Ruisi, G.; Barbieri, R. Synthesis and characterisation of diorganotin(IV) derivatives of 2-mercaptopyridine and crystal structure of diphenyl pyridine-2-thiolatochlorotin(IV). Appl. Organomet. Chem.1994, 8, 397–407.Search in Google Scholar
Sheldrick, G. M. SHELXL97. A computer program for crystal structure refinement. Version WinGX©, 1993–1997, Release 97-2. 1997a. University Göttingen, Germany.Search in Google Scholar
Sheldrick, G. M. SHELXS-97, A computer program for the solution of crystal structures. Version WinGX©, 1986–1997, Release 97-2. 1997b. University Göttingen, Germany.Search in Google Scholar
Sousa-Pedrares, A.; Casanova, M. I.; García-Vázquez, J. A.; Durán, M. L.; Romero, J.; Sousa, A.; Silver, J.; Titler, P. J. Synthesis and X-ray structures of tin(IV) and lead(II) complexes with heterocyclic thiones. Eur. J. Inorg. Chem.2003, 678–686.10.1002/ejic.200390094Search in Google Scholar
Thomas, B.; Paasch, S.; Steuernagel, S.; Eichele, K. Residual 31P, 35,37Cl dipolar coupling in 31P MAS spectra of chlorocyclophosphazenes. Solid State Nucl. Mag. Res.2001, 20, 108–117.Search in Google Scholar
Truflandier, L. A.; Brendler, E.; Wagler, J.; Autschbach, J. 29Si DFT/NMR observation of spin-orbit effect in metallasilatrane sheds some light on the strength of the metal→Si interaction. Angew. Chem. Int. Ed.2011, 50, 255–259.Search in Google Scholar
Valle, G.; Ettorre, R.; Vetorri, U.; Peruzzo, V.; Plazogna, G. Crystal structure and mass spectrometry of dichlorodimethylbis[2(1H)-pyridinethione-S]tin(IV). J. Chem. Soc., Dalton Trans.1987, 815–817.10.1039/dt9870000815Search in Google Scholar
Wächtler, E.; Kroke, E.; Wagler, J. Somewhere in between: the dichotomy E(II)→Pd(II) and E(IV)←Pd(0) (E=Ge,Sn) in 2-mercaptopyridyl bridged heterodinuclear complexes. 14th International Conference on the Coordination and Organometallic Chemistry of Germanium, Tin and Lead; Baddeck, Nova Scotia, Canada, July 14–19, 2013. Abstract No. 61.Search in Google Scholar
Wagler, J.; Hill, A. F.; Heine, T. Palladastannatrane – a PdII→SnIV Dative Bond. Eur. J. Inorg. Chem.2008, 4225–4229.10.1002/ejic.200800518Search in Google Scholar
Wagler, J.; Brendler, E. Metallasilatranes: palladium(II) and platinum(II) as lone-pair donors to silicon(IV). Angew. Chem. Int. Ed.2010a, 49, 624–627.Search in Google Scholar
Wagler, J.; Brendler, E.; Langer, T.; Pöttgen, R.; Heine, T.; Zhechkov, L. Ylenes in the MII→SiIV (M=Si, Ge, Sn) coordination mode. Chem. Eur. J.2010b, 16, 13429–13434.Search in Google Scholar
Wagler, J.; Heine, T.; Hill, A. F. Poly(methimazolyl)silanes—syntheses and molecular structures. Organometallics2010c, 29, 5607–5613.10.1021/om1004989Search in Google Scholar
Wagler, J.; Brendler, E.; Heine, T.; Zhechkov, L. Disilicon complexes with two hexacoordinate Si atoms: paddlewheel-y with (ClN4)Si-Si(S4Cl) and (ClN2S2)Si-Si(S2N2Cl) skeletons. Chem. Eur. J.2013, 19, 14296–14303.Search in Google Scholar
Wu, L.-M.; Chen, L.; Dia, J.; Lin, P.; Du, W.-X.; Wu, X.-T. trans-Bis[ethane-1,2-dithiolato(2-)-S,S′]bis(2-mercaptopyridine-S)tin(IV). Acta Crystallogr.2000, C56, e382.Search in Google Scholar
©2013 by Walter de Gruyter Berlin Boston
This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Articles in the same Issue
- Masthead
- Masthead
- Review
- Structural characterization of heterometallic platinum complexes with non-transition metals. Part V: Heterooligo- and heteropolynuclear complexes
- Research Articles
- {Sn10Si(SiMe3)2[Si(SiMe3)3]4}2-: cluster enlargement via degradation of labile ligands
- Unusual reaction pathways of gallium(III) silylamide complexes
- Molecular structures of pyridinethiolato complexes of Sn(II), Sn(IV), Ge(IV), and Si(IV)
- Arylphosphonic acid esters as bridging ligands in coordination polymers of bismuth
- Synthesis of germanium dioxide nanoparticles in benzyl alcohols – a comparison
- X-ray crystal structures of [(Cy2NH2)]3[C6H3(CO2)3]·4H2O and [i-Bu2NH2][(Me3 SnO2C)2C6H3CO2]
- Crystal and molecular structure of bis(di-n-propylammonium) dioxalatodiphenylstannate, [n-Pr2NH2]2[(C2O4)2SnPh2]
- Short Communication
- Synthesis and structural characterization of a dimeric N,N-dimethylformamide solvate of isobutyltin(IV) dichloride hydroxide, [iBuSnCl2(OH)(DMF)]2
Articles in the same Issue
- Masthead
- Masthead
- Review
- Structural characterization of heterometallic platinum complexes with non-transition metals. Part V: Heterooligo- and heteropolynuclear complexes
- Research Articles
- {Sn10Si(SiMe3)2[Si(SiMe3)3]4}2-: cluster enlargement via degradation of labile ligands
- Unusual reaction pathways of gallium(III) silylamide complexes
- Molecular structures of pyridinethiolato complexes of Sn(II), Sn(IV), Ge(IV), and Si(IV)
- Arylphosphonic acid esters as bridging ligands in coordination polymers of bismuth
- Synthesis of germanium dioxide nanoparticles in benzyl alcohols – a comparison
- X-ray crystal structures of [(Cy2NH2)]3[C6H3(CO2)3]·4H2O and [i-Bu2NH2][(Me3 SnO2C)2C6H3CO2]
- Crystal and molecular structure of bis(di-n-propylammonium) dioxalatodiphenylstannate, [n-Pr2NH2]2[(C2O4)2SnPh2]
- Short Communication
- Synthesis and structural characterization of a dimeric N,N-dimethylformamide solvate of isobutyltin(IV) dichloride hydroxide, [iBuSnCl2(OH)(DMF)]2