Abstract
Six tin complexes of the type (O,N,N,O)Sn and (O,N,N,O)SnCl2 (L1Sn, L2Sn, L3Sn, L1SnCl2, L2SnCl2, L3SnCl2) with di-anionic salen type tetradentate O,N,N,O chelators were synthesized and characterized by 1H, 13C, 119Sn NMR spectroscopy, single-crystal X-ray diffraction, and elemental analysis (L1Sn, L2Sn, L1SnCl2, L2SnCl2, L3SnCl2) or 119Sn solid state NMR spectroscopy and elemental analysis (L3Sn). The length of the oligomethylene bridge of the ligands (L1)2-, (L2)2- and (L3)2- (which are the anions of di-, tri- and tetramethylene-α,ω-N,N′-bis(2-hydroxyacetophenoneimine, respectively) was found to have significant influence on the configuration of the tin coordination sphere. In detail, the Sn atom in L1Sn occupies the apex of a (O,N,N,O) square pyramid, whereas L2Sn resembles a see-saw setup of the tin coordination sphere, and L3Sn appears to be of oligomeric nature concluded from its distinctly lower solubility and different 119Sn NMR shift. In the monomolecular Sn(IV) compounds the Cl atoms are trans-disposed in L1SnCl2, and cis-arrangements of the Cl atoms in combination with a mer-fac-arrangements of the (O,N,N,O) ligands are found in L2SnCl2 and L3SnCl2.
Introduction
The coordination chemistry of dianionic salen-type tetradentate chelators [salen=ethylene-N,N′-bis(salicylideneimine)] has been explored with group 14 elements in the past decades. An early report of a crystal structure of (salen)SnMe2 (Calligaris et al., 1972) indicated trans-configuration of the monodentate substituents and mer-mer-configuration of the tetradentate chelator in such tin complexes. In the course of our investigations of mainly silicon complexes and, to a lesser extent, germanium and tin complexes of salen-type chelators, we noticed some remarkable features that distinguish such silicon complexes from related tin compounds, as highlighted in Scheme 1. Whereas crystallographic evidence for hexacoordinate Si salen complexes with two monodentate substituents (Mucha et al., 1998, 1999; Wagler et al., 2004, 2005a,b,c, 2009; González-García et al., 2009) points at the exclusive formation of complexes with trans-disposed substituents (like in I and the germanium analog II, Wagler et al., 2005a, 2009), the related tin compound III was found to prefer cis-arrangement of the monodentate substituents (Wagler et al., 2009). In the latter case, the salen-type ligand is mer-fac-arranged, and the same arrangement of the tetradentate ligand had been reported for a salen-tin(IV) complex with cis-forced substituents within a stannacycle (Agustin et al., 1999a). In a related reaction with a ligand that bears imine-α-hydrogen atoms, the Sn complex VI (Wagler et al., 2005b) even resists the formation of an enamine like in Si and Ge compounds IV and V (Wagler et al., 2002, 2005b) in favor of the formation of a mer-fac-configured hexacoordinate Sn complex. In Si(IV) complexes this mer-fac-arrangement of the tetradentate ligand can be enforced by the steric constraints of silacycles, e.g., in VII (Wagler et al., 2005a) and related compounds (Wagler et al., 2006; Wagler and Roewer, 2007). These results and literature reports of mer-mer-configured hexacoordinate Sn(IV) salen complexes, such as group VIII of complexes (Jing et al., 2004; Darensbourg et al., 2005) indicate enhanced flexibility of the salen ligand in the tin coordination sphere. This flexibility of the tetradentate ligand appears to be founded on the ethylene bridge, because hexacoordinate silicon (Lippe et al., 2009) and tin complexes (Teoh et al., 1997; Dey et al., 1999a,b; Yearwood et al., 2002; Jing et al., 2004; Darensbourg et al., 2005; Wang et al., 2010) with two monodentate substituents and related but more rigid ligand backbones (e.g., 1,2-phenylene or 1,2-cyclohexanediyl bridged Schiff base ligands) have been found in trans form exclusively in terms of crystallographically proven isomers. Furthermore, salen-tin complexes already have been reported for Sn(II), like complex IX (Kuchta et al., 1999; Westerhausen et al., 2003; Jing et al., 2004), the latter of which also have been utilized as ligands in transition metal coordination chemistry (Agustin et al., 2000a,b), silylene analogs (i.e., Si(II) complexes) are still unknown.

Selected Si, Ge, and Sn complexes with salen-type ligands.
A search in the Cambridge Structural Database, CSD (ConQuest 1.16, 2013) revealed a noticeable lack of crystallographically characterized tin complexes with salen-type (ONNO) ligands, which bear more than two methylene groups in the central (N-Sn-N) chelate. The convenient availability of a series of such salen-type ligands, H2L1, H2L2, H2L3, which has been used in one of our previous studies of silicon complexes (Wagler et al., 2005b), the general feasibility of syntheses of tin(II) and tin(IV) complexes with salen-type ligands, and the enhanced coordinative flexibility of tin (cf. silicon) served as motivation for the herein presented study.
Results and discussion
As shown in Scheme 2, the reaction of bis(bis(trimethylsilyl)amino)stannylene with the tetradentate Schiff base ligand (H2L1, H2L2 or H2L3) affords the corresponding tin(II) complex (L1Sn, L2Sn or L3Sn). Interestingly, compounds L1Sn and L2Sn were soluble in chloroform and could be characterized by NMR spectroscopy in solution state, whereas compound L3Sn already precipitated from the reaction mixture (in THF) and could not be dissolved in chloroform or DMSO. Thus, we have not succeeded in crystallizing L3Sn, and this complex was characterized by solid state 119Sn NMR only. The 119Sn NMR shifts (L1Sn, CDCl3: -571 ppm; L2Sn, CDCl3: -620 ppm; L3Sn, solid: -524 ppm) are in the expected range for tin(II) complexes with salen-type ligands [e.g., -536 ppm (Westerhausen et al., 2003), -501, -517, -522, -527 ppm (Jing et al., 2004)] and other Sn(II) complexes with four donor atoms, such as aryloxy-O or sp2 (amide, imine) N atoms; for comparison with 119Sn NMR shifts of other Sn(II) complexes with such donor spheres, see Scheme 3 (IX: van den Bergen et al., 1990; X: Brendler et al., 2011). The 1H and 13C NMR data of L1Sn and L2Sn suggest symmetric monomeric complexes in chloroform solution. In both cases the spectra show a single set of signals in accord with a Cs symmetric molecule. In the solid state structure (as determined by single-crystal X-ray diffraction, Figure 1, Table 1) compound L1Sn is monomeric, and the Sn atom is located at the apex of an idealized square pyramid with ONNO base, thus this structure accords well with the proposed Cs symmetric monomers of L1Sn in solution and similar to the molecular structures of previously published tetracoordinate Sn(II) complexes with related (ONNO) ligand systems (Atwood et al., 1995; Agustin et al., 1999b; Kuchta et al., 1999; Westerhausen et al., 2003; Jing et al., 2004) and monomeric Sn(II) compounds with other (ONNO) ligand systems (Zöller et al., 2011). The molecular structure of L2Sn in the solid state is surprisingly different (Figure 1). The tetradentate ligand furnishes a see-saw-type coordination sphere about the Sn atom (with O1 and N2 in apical positions and O2 and N1 on equatorial sites). Furthermore, each of the apical O1 atoms establishes a contact to the tin atom of a neighboring molecule, thus forming dimers with 4+1 coordinated tin atoms (Scheme 4). The highly unsymmetric Sn1-O1···Sn1* bridges (Sn-O separations are 2.15 and 3.04 Å, respectively) already hint at weak bridging and possible dissociation into monomers in solution. In the solid state structure of L2Sn the tetradentate ligand adopts a mer-fac configuration (mer-O1,N1,N2; fac-N1,N2,O2) with respect to an idealized octahedral coordination sphere (the additional two coordination sites being occupied by the O1* donor and the Sn-located lone pair). This feature relates the molecular conformation of L2Sn to that of the tin(IV) complex L2SnCl2 (vide infra). As both the formation of dimers (Berends et al., 2009; Iovkova-Berends et al., 2011, 2012a,b) and polymers (Zöller et al., 2013) via Sn2 O2 four-membered cycles has been encountered with stanna(II)-ocanes, we attribute the poor solubility of L3Sn to a similar kind of formation of polymeric networks.

Generic routes of the syntheses of compounds L1-3Sn and L1-3SnCl2.

Examples of tetracoordinate Sn(II) complexes and their 119Sn NMR shifts.
![Figure 1 Molecular structures of (from top) L1Sn and [L2Sn]2 in the solid state. 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. For L1Sn, only the predominant position (ca. 70% occupancy) of the cross-wise disordered ethylene bridge is depicted. The dimer [L2Sn]2 is located around a center of inversion (operation 1-x, 1-y, -z), which generates the symmetry equivalent atoms (asterisked labels) from the asymmetric unit. Selected bond lengths (Å) and angles (°) for L1Sn: Sn1-O1 2.1093(14), Sn1-O2 2.1032(14), Sn1-N1 2.3554(18), Sn1-N2 2.3511(16), O1-C1 1.331(2), O2-C10 1.321(2), N1-C7 1.295(3), N2-C16 1.293(3), O1-Sn1-O2 76.43(5), O1-Sn1-N1 76.19(6), O2-Sn1-N2 77.43(5), N1-Sn1-N2 70.74(6), O1-Sn1-N2 113.67(6), O2-Sn1-N1 124.20(6); for [L2Sn]2: Sn1-O1 2.1524(8), Sn1-O2 2.0944(8), Sn1···O1* 3.035(1), Sn1-N1 2.2906(9), Sn1-N2 2.5316(9), O1-C1 1.324(1), O2-C10 1.325(1), N1-C7 1.304(1), N2-C16 1.290(1), O1-Sn1-O2 82.86(3), O1-Sn1-N1 79.59(3), O2-Sn1-N2 74.78(3), N1-Sn1-N2 74.24(3), O1-Sn1-N2 143.70(3), O2-Sn1-N1 94.65(3).](/document/doi/10.1515/mgmc-2014-0004/asset/graphic/mgmc-2014-0004_fig1.jpg)
Molecular structures of (from top) L1Sn and [L2Sn]2 in the solid state. 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. For L1Sn, only the predominant position (ca. 70% occupancy) of the cross-wise disordered ethylene bridge is depicted. The dimer [L2Sn]2 is located around a center of inversion (operation 1-x, 1-y, -z), which generates the symmetry equivalent atoms (asterisked labels) from the asymmetric unit. Selected bond lengths (Å) and angles (°) for L1Sn: Sn1-O1 2.1093(14), Sn1-O2 2.1032(14), Sn1-N1 2.3554(18), Sn1-N2 2.3511(16), O1-C1 1.331(2), O2-C10 1.321(2), N1-C7 1.295(3), N2-C16 1.293(3), O1-Sn1-O2 76.43(5), O1-Sn1-N1 76.19(6), O2-Sn1-N2 77.43(5), N1-Sn1-N2 70.74(6), O1-Sn1-N2 113.67(6), O2-Sn1-N1 124.20(6); for [L2Sn]2: Sn1-O1 2.1524(8), Sn1-O2 2.0944(8), Sn1···O1* 3.035(1), Sn1-N1 2.2906(9), Sn1-N2 2.5316(9), O1-C1 1.324(1), O2-C10 1.325(1), N1-C7 1.304(1), N2-C16 1.290(1), O1-Sn1-O2 82.86(3), O1-Sn1-N1 79.59(3), O2-Sn1-N2 74.78(3), N1-Sn1-N2 74.24(3), O1-Sn1-N2 143.70(3), O2-Sn1-N1 94.65(3).

Steric representation of the conformations of L1Sn (monomeric, square pyramidal Sn coordination sphere) and L2Sn (dimeric, see-saw Sn coordination of each monomeric unit).
Parameters of data collection and structure refinement for L1Sn, L2Sn, L1SnCl2, L2SnCl2 and L3SnCl2.
L1Sn | L2Sn | L1SnCl2 | L2SnCl2 | L3SnCl2 | |
---|---|---|---|---|---|
Empirical formula | C18 H18 N2 O2 Sn | C19 H20 N2 O2 Sn | C18 H18 Cl2 N2 O2 Sn | C19 H20 Cl2 N2 O2 Sn | C20 H22 Cl2 N2 O2 Sn |
Formula weight | 413.03 | 427.06 | 483.93 | 497.96 | 511.99 |
T (K) | 150(2) | 150(2) | 200(2) | 150(2) | 200(2) |
λ (Å) | 0.71073 | ||||
Crystal system | Orthorhombic | Triclinic | Monoclinic | Monoclinic | Monoclinic |
Space group | P21 21 21 | P-1 | P21/c | P21 | P21 |
Unit cell dimensions | |||||
a (Å) | 6.6612(3) | 8.1851(5) | 16.0521(5) | 7.8351(4) | 8.3131(5) |
b (Å) | 14.4974(5) | 9.1013(6) | 12.5497(3) | 11.5687(8) | 12.6454(11) |
c (Å) | 16.9498(8) | 11.8043(8) | 18.8112(6) | 21.3757(11) | 10.0364(7) |
α (°) | 90 | 81.854(5) | 90 | 90 | 90 |
β (°) | 90 | 84.661(5) | 105.442(2) | 93.674(4) | 110.308(5) |
γ (°) | 90 | 81.684(5) | 90 | 90 | 90 |
V (Å3) | 1636.84(12) | 858.97(10) | 3652.70(18) | 1933.55(19) | 989.47(13) |
Z/Dc (g/cm3) | 4/1.68 | 2/1.65 | 8/1.76 | 4/1.71 | 2/1.72 |
μ (mm-1) | 1.6 | 1.5 | 1.7 | 1.6 | 1.6 |
F(000) | 824 | 428 | 1920 | 992 | 512 |
Crystal size (mm) | 0.25×0.15×0.10 | 0.50×0.35×0.15 | 0.40×0.20×0.10 | 0.40×0.03×0.02 | 0.25×0.20×0.18 |
θ range for data collection | 2.8–30.0 | 2.7–32.0 | 2.5–32.0 | 2.6–26.0 | 2.6–30.0 |
Reflections collected | 27574 | 18158 | 62649 | 16788 | 10889 |
Independent reflections/R(int) | 4780/0.0305 | 5972/0.0214 | 12673/0.0366 | 7463/0.0782 | 5752/0.0197 |
Completeness to θmax | 99.9% | 99.9% | 100% | 99.9% | 100% |
Refinement | Full matrix least-squares on F2 | ||||
Data/restraints/parameters | 4780/3/229 | 5972/0/219 | 12673/0/455 | 7463/1/481 | 5752/1/246 |
Goodness-of-fit on F2 | 1.072 | 1.090 | 1.069 | 1.005 | 1.087 |
χ(Flack) | -0.031(14) | 0.0(4)a | -0.018(13) | ||
R1/wR2 [I>2σ(I)] | 0.0177/0.0392 | 0.0168/0.0435 | 0.0235/0.0524 | 0.0454/0.0881 | 0.0191/0.0474 |
R1/wR2 (all data) | 0.0193/0.0398 | 0.0180/0.0441 | 0.0302/0.0547 | 0.0758/0.0990 | 0.0209/0.0483 |
Largest diff. peak and hole, eÅ-3 | 0.653/-0.633 | 0.542/-0.359 | 0.753/-0.569 | 0.901/-1.049 | 0.498/-0.252 |
aThis structure was refined as a racemic twin. Before twin refinement the Flack parameter was 0.53(3).
As outlined in Scheme 2, the tin(IV) complexes L1SnCl2, L2SnCl2 and L3SnCl2 were synthesized from the respective ligand and tin(IV) chloride by base supported HCl elimination (using triethylamine as the sacrificial base). Their 119Sn NMR shifts in DMSO (L1SnCl2, -610 ppm; L2SnCl2, -623 ppm; L3SnCl2, -617 ppm) are characteristic for hexacoordinate (salen)SnCl2 complexes (shifts of -594 and -596 ppm have been reported by Jing et al., 2004, for CD2 Cl2 solutions of related complexes), and they appear in a narrow range in spite of the different chelate sizes and ligand conformations. According to 1H and 13C NMR spectroscopy compound L1SnCl2 exhibits two chemically equivalent halves of the salen ligand in solution, and this is in accord with the molecular configuration observed in the solid state (Figure 2). Compounds L2SnCl2 and L3SnCl2 each produce 1H and 13C NMR spectra characteristic of salen ligands with non-equivalent half-sides. Apparently, in solution these complexes retain the molecular conformation that is found in their solid state structures (Figure 2). In both cases the tetradentate ligand adopts mer-fac configuration. Retention of this configuration in solution is in contrast to the observation made with compounds III and VI (Wagler et al., 2005b, 2009), which produce solution state 1H and 13C NMR spectra characteristic of complexes with chemically equivalent ligand halves and thus indicate formation of the mer-mer configuration or rapid configurational exchange in solution.

Molecular structures of (from top) L1SnCl2, L2SnCl2 and L3SnCl2 in the solid state. 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. For L1SnCl2 and L2SnCl2, which contain two molecules of the same isomer in the asymmetric unit each, only one molecule is depicted as a representative example. Selected bond lengths (Å) for L1SnCl2: Sn1-Cl1 2.4518(4), Sn1-Cl2 2.4179(4), Sn1-O1 2.0026(11), Sn1-O2 2.0054(11), Sn1-N1 2.1525(13), Sn1-N2 2.1676(12), O1-C1 1.345(2), O2-C10 1.348(2), N1-C7 1.298(2), N2-C16 1.303(2), O1-Sn1-O2 103.70(5), O1-Sn1-N1 87.69(5), O2-Sn1-N2 88.57(5), N1-Sn1-N2 80.12(5), O1-Sn1-N2 167.71(5), O2-Sn1-N1 167.99(5), Cl1-Sn1-Cl2 175.32(2); for L2SnCl2: Sn1-Cl1 2.389(2), Sn1-Cl2 2.402(2), Sn1-O1 1.995(5), Sn1-O2 2.021(6), Sn1-N1 2.211(7), Sn1-N2 2.196(7), O1-C1 1.362(10), O2-C9 1.364(10), N1-C7 1.296(10), N2-C15 1.302(10), O1-Sn1-O2 89.4(2), O1-Sn1-N1 83.2(2), O2-Sn1-N2 82.5(2), N1-Sn1-N2 82.5(2), O1-Sn1-N2 163.8(2), O2-Sn1-N1 88.1(2), Cl1-Sn1-Cl2 91.27(8), Cl1-Sn1-O2 174.93(17), Cl2-Sn1-N1 178.54(18); for L3SnCl2: Sn1-Cl1 2.4187(5), Sn1-Cl2 2.4083(5), Sn1-O1 2.0228(14), Sn1-O2 2.0205(15), Sn1-N1 2.1990(17), Sn1-N2 2.2073(18), O1-C1 1.340(2), O2-C9 1.338(2), N1-C7 1.311(3), N2-C15 1.307(3), O1-Sn1-O2 91.49(8), O1-Sn1-N1 82.90(8), O2-Sn1-N2 82.87(7), N1-Sn1-N2 86.95(7), O1-Sn1-N2 87.96(6), O2-Sn1-N1 168.55(7), Cl1-Sn1-Cl2 90.09(2), Cl1-Sn1-O1 174.15(7), Cl2-Sn1-N2 174.66(5).
From the structural and spectroscopic observations made with compounds L1SnCl2, L2SnCl2, and L3SnCl2, we conclude that upon transition from the ethylene bridged ligand L1 to the more flexible propylene bridged ligand L2, the mer-fac configuration of the tetradentate chelator is favored over the mer-mer configuration. To prove this hypothesis, we have performed quantum chemical analyses (with Gaussian 09, Frisch et al., 2009) of the relative energies of the mer-mer, mer-fac, and fac-fac configurational isomers of L1SnCl2 and L2SnCl2 (Scheme 5, Figure 3). In addition to the evaluation of the relative energies obtained by optimization of the molecular structures at the DFT-B3LYP (CHNOCl: 6-31G(d); Sn: SDD) level the trends were confirmed by single-point energy analyses at the RHF-MP2 (CHNOCl: 6-31G(d); Sn: SDD) level. In case of L1SnCl2 the molecule with mer-mer configuration should indeed represent the stable isomer, but the mer-fac isomer is less favorable by only 1–2 kcal/mol, whereas formation of the fac-fac isomer would require significantly more energetic effort (7–10 kcal/mol). These results are in agreement with the occasional observation of mer-fac configured (salen)Sn(IV) complexes (compounds III and VI) in the solid state (most likely driven by packing effects) and their transformation into the mer-mer isomers in solution. Noteworthy, for related silicon complexes with ethylene bridged tetradentate ONNO ligand, the formation of the mer-fac and the fac-fac isomer out of the mer-mer isomer would require 5.6 and 18.7 kcal/mol, respectively (Seiler et al., 2005). Thus, the tin(IV) complexes appear to be more flexible than their Si(IV) congeners, which is in agreement with the exclusive observation of mer-mer isomers of (salen)Si(IV) complexes with two monodentate substituents as pointed out earlier. For L2SnCl2 (Figure 3) the mer-mer isomer is indeed less stable than the mer-fac isomer and, to our surprise, the mer-fac and fac-fac isomer exhibit similar energetics. The lower stability of the mer-mer isomer is in accord with our NMR spectroscopic observation of the retention of the mer-fac configuration in solution. Furthermore, the 1H and 13C NMR data do not hint at transformation into the fac-fac isomer in solution (which would have to give rise to a set of signals characteristic of chemically equivalent ligand halves because of their relation via a two-fold axis).

Principle isomers of complexes of the type LnSnCl2 with hexacoordinate Sn atom. The notations mer-mer, mer-fac, fac-fac refer to the relative positions of three donor atoms of the tetradentate (ONNO) ligand, i.e., (ONN) sequence-(NNO) sequence.

Molecular structures of (from left) the mer-mer, mer-fac and fac-fac isomer of L1SnCl2 (top row) and L2SnCl2 (bottom row), optimized with Gaussian09 (Frisch et al., 2009) at the B3LYP (CHNOCl: 6-31G(d); Sn: SDD) level. For each row, the relative energies of the isomers (in kcal/mol) are given below the molecular structure. The relative energies in parentheses (also in kcal/mol) were obtained from single-point analyses at the MP2 (CHNOCl: 6-31G(d); Sn: SDD) level.
Conclusions
The transition from ethylene bridged salen-type ligands (such as L1) to related tri- or tetramethylene bridged ligands (such as L2 or L3) has significant effect on the molecular conformations of tin(II) and tin(IV) complexes of these tetradentate (ONNO) chelators. Whereas for L1 a mer-mer arrangement generally is favorable (as the base of a square pyramidal coordination sphere or within the equatorial plane of an octahedral coordination sphere) and transformation into the mer-fac configuration energetically is less favorable in solution but still possible for the purpose of crystallization, the more flexible ligands L2 and L3 shift the preferences to the mer-fac and (as calculated) fac-fac configuration. Even though the latter has not been observed with that kind of tin complexes, cobalt(III) complexes of the type [LCo(pyridine)2]+ have been found to form the mer-mer isomer for L=L2 and the fac-fac isomer for L=L3 (Schenk et al., 2007). Thus, we predict that the fac-fac coordination mode of a tetradentate salen-type Schiff base ligand in the coordination sphere of tin also is very likely to be observed in case of (ONNO) ligands with tri- or tetramethylene bridge.
Experimental
Ligands H2L1, H2L2 and H2L3 were prepared according to a method published previously (Wagler et al., 2005b). Bis(bis(trimethylsilyl)amino)stannylene, which also is known from the literature (Fjeldberg et al., 1983) was prepared from commercially available sodium bis(trimethylsilyl)amide solution (2M in THF, Sigma-Aldrich, Steinheim, Germany) and SnCl2(dioxane) (Morrison et al., 1967). All other chemicals used were commercially available and used without further purification. Tetrahydrofuran (THF) and triethylamine were distilled from sodium/benzophenone and stored under dry argon. Dichloromethane (DCM) was distilled from calcium hydride and stored under argon. Acetonitrile, dimethyl sulfoxide (DMSO) and chloroform (stabilized with amylenes) were stored over molecular sieves 3 Å. Elemental microanalyses (C, H, N) were performed on a ‘Vario Micro Cube’ analyzer (Elementar, Hanau, Germany). NMR spectra of solutions were recorded on an Avance 500 spectrometer (Bruker Biospin, Rheinstetten, Germany) and internally referenced to tetramethylsilane for 1H and 13C and externally referenced to tetramethylstannane for 119Sn. The 119Sn solid state NMR spectrum of L3Sn was recorded on an Avance 400 WB spectrometer (Bruker Biospin, Rheinstetten, Germany) using a 4-mm zirconia (ZrO2) spinner. Single-crystal X-ray diffraction data sets were collected on an IPDS-2(T) diffractometer (Stoe, Darmstadt, Germany) using Mo Kα-radiation. 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). Hydrogen atoms were placed in idealized positions and refined isotropically (riding model). The structure of L1Sn bears a cross-wise disorder of the ligand ethylene bridge with site occupancies of 0.702(7) and 0.298(7). The structure of L2SnCl2 was refined as a racemic twin with twin populations of 0.54(3) and 0.46(3), and the Sn atoms of this structure were refined in two alternative positions with site occupancies of 0.949(1) and 0.051(1), the latter probably arising from stacking faults in the crystal lattice, thus simulating disorder positions. The remaining atoms (C,H,N,O,Cl) of those phantom molecules have not been detected because of their relatively low occupancy, and therefore the ligand atoms have been refined without disorder (with full site occupancy). Crystallographic data for the structures reported in this paper have been deposited with the Cambridge Crystallographic Data Center as supplementary material publication no. CCDC-990379 (L1Sn), CCDC-990380 (L2Sn), CCDC-990381 (L1SnCl2), CCDC-990382 (L2SnCl2), and CCDC-990383 (L3SnCl2). 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 L1Sn
To a stirred suspension of ligand H2L1 (4.00 g, 13.5 mmol) in THF (75 mL) at room temperature was added Sn(N(SiMe3)2)2 (5.61 g, 13.5 mmol), which had been heated to 60°C (to lower its viscosity). Upon stirring at room temperature for 10 min a clear orange solution was obtained, wherefrom the product crystallized upon standing at room temperature for 2 days (crystals were suitable for X-ray diffraction analysis). Thereafter, the product was filtered and washed with THF (2 mL) and dried in vacuo. Yield 2.99 g (7.24 mmol, 54%) of yellow crystalline solid that decomposes (in a sealed glass capillary) at 257°C without melting. 1H NMR (500.1 MHz, CDCl3) δ: 2.40 (s, 6H, CH3), 3.70–3.92 (m, 4H, NCH2 CH2 N), 6.65 (m, 2H, aryl CH), 7.09 (d, 3J(1H-1H)=8.1 Hz, 2H, aryl CH), 7.24 (m, 2H, aryl CH), 7.36 (d, 3J(1H-1H)=8.1 Hz, 2H, aryl CH). 13C{1H} NMR (125.8 MHz, CDCl3) δ: 19.1 (CH3), 49.7 (NCH2CH2 N), 116.1, 124.1, 124.8, 129.2, 133.0 (aryl C), 164.2 (aryl C–O), 172.8 (C=N). 119Sn NMR (186.5 MHz, CDCl3) δ: -571. Elemental analyses: calculated (%) for C18 H18 N2 O2 Sn (413.03 g/mol): C 52.34, H 4.39, N 6.78, found: C 51.9, H 4.4, N 7.1.
Synthesis of L2Sn
To a stirred solution of ligand H2L2 (4.00 g, 12.9 mmol) in THF (75 mL) at room temperature was added Sn(N(SiMe3)2)2 (5.44 g, 13.1 mmol) that had been heated to 60°C (to lower its viscosity). Upon stirring at room temperature for 1 min the clear orange solution thus obtained was stored in a refrigerator (7°C) for 5 days to afford a crystalline product with crystals suitable for X-ray diffraction analysis. Thereafter, the product was filtered and washed with THF (3 mL) and dried in vacuo. Yield 2.86 g (6.70 mmol, 52%) of yellow crystalline soild, m.p. 207°C (sealed glass capillary). 1H NMR (500.1 MHz, CDCl3) δ: 1.95 (m, 1H, NCH2 CH2 CH2 N), 2.36 (s, 6H, CH3), 2.87 (m, 1H, NCH2 CH2 CH2 N), 3.88 (m, 4H, NCH2 CH2 CH2 N), 6.59 (t, 3J(1H-1H)=7.8 Hz, 2H, aryl CH), 6.83 (d, 3J(1H-1H)=7.8 Hz, 2H, aryl CH), 7.18 (dt, 3J(1H-1H)=7.8 Hz, 4J(1H-1H)=1.4 Hz, 2H, aryl CH), 7.34 (dd, 3J(1H-1H)=7.8 Hz, 4J(1H-1H)=1.4 Hz, 2H, aryl CH). 13C{1H} NMR (125.8 MHz, CDCl3) δ: 18.7 (CH3), 27.6 (NCH2CH2 CH2 N), 51.4 (NCH2 CH2CH2 N), 115.8, 124.1, 124.2, 129.4, 132.9 (aryl C), 164.3 (aryl C–O), 171.9 (C=N). 119Sn NMR (186.5 MHz, CDCl3) δ: -620. Elemental analyses: calculated (%) for C19 H20 N2 O2 Sn (427.06 g/mol): C 53.43, H 4.72, N 6.56, found: C 53.1, H 4.7, N 6.4.
Synthesis of L3Sn
Upon initial experiments similar to the synthesis of L1Sn, which did afford fine precipitates of very poor solubility, the synthesis was repeated using lower concentrations of the starting materials but also failed to give single-crystals suitable for X-ray diffractometry. To a stirred suspension of ligand H2L3 (0.44 g, 1.36 mmol) in THF (35 mL) at room temperature was added Sn(N(SiMe3)2)2 (0.57 g, 1.37 mmol) that had been heated to 60°C (to lower its viscosity). Upon stirring at room temperature for some minutes a clear orange solution was obtained, wherefrom the product precipitated as a fine powder upon standing at room temperature. After 8 days, the product was filtered and washed with THF (2 mL) and dried in vacuo. Yield 0.32 g (0.72 mmol, 53%) of a fine yellow powder. This compound has very poor solubility in solvents, such as chloroform, dichloromethane, toluene, THF, and DMSO, and therefore we have not been able to record solution NMR spectra (neither 119Sn and 13C nor 1H). 119Sn NMR (149.2 MHz, solid state) δ: -524. Elemental analyses: calculated (%) for C20 H22 N2 O2 Sn (441.11 g/mol): C 54.46, H 5.03, N 6.35, found: C 54.5, H 5.3, N 6.3.
Synthesis of L1SnCl2
At room temperature a suspension of ligand H2L1 (2.10 g, 7.09 mmol) and triethylamine (1.38 g, 13.6 mmol) in DCM (40 mL) was added to a stirred solution of SnCl4 (1.78 g, 6.83 mmol) in DCM (20 mL). The light yellow suspension formed was stirred at room temperature for 3 h, whereupon the white product was filtered, washed with DCM (2×2 mL) and dried in vacuo. This crude product was then recrystallized from DMSO (to yield crystals suitable for X-ray diffraction analysis), filtered off, washed with DCM (8 mL) and dried in vacuo. Yield 2.10 g (4.34 mmol, 61%) of colorless crystalline solid, which decomposes (in a sealed glass capillary) at 301°C without melting. 1H NMR (500.1 MHz, DMSO-D6) δ: 2.74 (s, 6H, CH3), 4.21 (s, Sn satellites of 3J(1H-117/119Sn)=33.4 Hz, 117Sn and 119Sn not resolved, 4H, NCH2 CH2 N), 6.84–6.94 (m, 4H, aryl CH), 7.43 (m, 2H, aryl CH), 7.81 (dd, 3J(1H-1H)=4.1 Hz, 4J(1H-1H)=1.4 Hz, 2H, aryl CH). 13C{1H} NMR (125.8 MHz, DMSO-D6) δ: 21.1 (CH3), 47.1 (NCH2CH2 N), 118.7, 120.6, 123.3, 132.1, 135.3 (aryl C), 163.3 (aryl C–O), 179.8 (C=N). 119Sn NMR (186.5 MHz, DMSO-D6) δ: -610. Elemental analyses: calculated (%) for C18 H18 N2 O2 SnCl2 (483.93 g/mol): C 44.67, H 3.75, N 5.79, found: C 44.3, H 3.8, N 5.7.
Synthesis of L2SnCl2
At room temperature a solution of ligand H2L2 (1.00 g, 3.22 mmol) and triethylamine (0.65 g, 6.42 mmol) in DCM (20 mL) was added to a stirred solution of SnCl4 (0.84 g, 3.22 mmol) in DCM (10 mL). The resulting yellow solution was stirred at room temperature for 10 min and then stored at 7°C overnight, whereupon colorless crystals of Et3 NHCl formed. The hydrochloride was then filtered off and the filtrate was stored at 7°C for 5 days to afford the product as a white precipitate, which was filtered off, washed with DCM (2×2.5 mL) and dried in vacuo (0.68 g crude product). This crude product was then recrystallized from acetonitrile (28 mL) to yield crystals suitable for X-ray diffraction analysis. After crystallization had commenced, the product solution was stored at -24°C for 1 week, then the crystalline product was filtered off, washed with acetonitrile (1 mL) and dried in vacuo. Yield 0.22 g (0.44 mmol, 14%) of colorless crystalline solid, which decomposes (in a sealed glass capillary) at 333°C without melting. 1H NMR (500.1 MHz, DMSO-D6) δ: 1.86–2.26 (m, 2H, NCH2 CH2 CH2 N), 2.69 (s, 3H, CH3), 2.71 (s, 3H, CH3), 3.91, 4.18, 4.27, 4.45 (4m, 4H, NCH2 CH2 CH2 N), 6.74–6.95 (m, 4H, aryl CH), 7.34–7.44 (m, 2H, aryl CH), 7.71–7.81 (m, 2H, aryl CH). 13C{1H} NMR (125.8 MHz, DMSO-D6) δ: 19.9, 20.8 (2CH3), 24.8 (NCH2CH2 CH2 N), 45.2, 45.5 (NCH2 CH2CH2 N), 118.6, 118.7, 121.6, 122.5, 122.9, 123.1, 131.0, 131.4, 134.5, 134.7 (aryl C), 163.8, 164.6 (aryl C–O), 181.0, 182.1 (C=N). 119Sn NMR (186.5 MHz, DMSO-D6) δ: -623. Elemental analyses: calculated (%) for C19 H20 N2 O2 SnCl2 (497.96 g/mol): C 45.82, H 4.05, N 5.63, found: C 45.4, H 4.2, N 5.7.
Synthesis of L3SnCl2
At room temperature a suspension of ligand H2L3 (1.01 g, 3.11 mmol) and triethylamine (0.63 g, 6.23 mmol) in DCM (20 mL) was added to a stirred solution of SnCl4 (0.81 g, 3.11 mmol) in DCM (10 mL). The yellow solution formed was stirred at room temperature for 2 h and then stored at -24°C, whereupon crystals of Et3 NHCl formed, which were filtered off. From the filtrate the volatiles were removed by vacuum condensation, and the solid residue was stirred in chloroform (20 mL). The white solid thus obtained was filtered off, washed with chloroform (2 mL) and dried in vacuo. This crude product was then recrystallized from DMSO (to yield crystals suitable for X-ray diffraction analysis), filtered off, washed with DCM (2×2 mL) and dried in vacuo. Yield 0.52 g (1.00 mmol, 32%) of colorless crystalline solid. 1H NMR (500.1 MHz, DMSO-D6) δ: 1.51–1.61 (m, 2H, NCH2 CH2 CH2 CH2 N), 1.81 (m, 1H, NCH2 CH2 CH2 CH2 N), 1.99 (m, 1H, NCH2 CH2 CH2 CH2 N), 2.23 (m, 1H, NCH2 CH2 CH2 CH2 N), 2.63 (s, 3H, CH3), 2.78 (s, 3H, CH3), 4.02–4.09 (m, 2H, NCH2 CH2 CH2 CH2 N), 4.42 (m, 1H, NCH2 CH2 CH2 CH2 N), 6.73–6.82 (m, 2H, aryl CH), 6.85–6.93 (m, 2H, aryl CH), 7.36–7.43 (m, 2H, aryl CH), 7.76 (d, 3J(1H-1H)=8.4 Hz, 1H, aryl CH), 7.83 (d, 3J(1H-1H)=8.9 Hz, 1H, aryl CH). 13C{1H} NMR (125.8 MHz, DMSO-D6) δ: 21.2, 21.4 (2CH3), 27.5, 29.3 (NCH2CH2CH2 CH2 N), 49.0, 49.8 (NCH2 CH2 CH2CH2 N), 118.4, 118.6, 122.1, 122.2, 122.7, 123.0, 131.1, 131.3, 134.5, 135.0 (aryl C), 163.0, 163.6 (aryl C–O), 181.9, 182.9 (C=N). 119Sn NMR (186.5 MHz, DMSO-D6) δ: -617. Elemental analyses: calculated (%) for C20 H22 N2 O2 SnCl2 (511.99 g/mol): C 46.92, H 4.33, N 5.47, found: C 46.4, H 4.4, N 5.5.
References
Agustin, D.; Rima, G.; Gornitzka, H.; Barrau, J. Stable heterocyclic (Schiff base) divalent group 14 element species M-O-Schiff base-O (M=Ge, Sn, Pb). J. Organomet. Chem. 1999a, 592, 1–10.Suche in Google Scholar
Agustin, D.; Rima, G.; Gornitzka, H.; Barrau, J. Tricoordinated species of group 14. (Schiff base) M = E (M = Ge, Sn; E = N-SiMe3, S, Se). Main Group Met. Chem. 1999b, 22, 703–712.Suche in Google Scholar
Agustin, D.; Rima, G.; Gornitzka, H.; Barrau, J. (Schiff base) Divalent group 14 element species: manganese and iron complexes (Salen)M=Mn(CO)2(η5-C5 H5) (M14 = Ge, Sn, Pb) and (Salen)Sn=Fe(CO)4. Inorg. Chem. 2000a, 39, 5492–5495.Suche in Google Scholar
Agustin, D.; Rima, G.; Gornitzka, H.; Barrau, J. Transition metal complexes of (Schiff base) divalent group 14 element species [(salen)M]n=M′(CO)6–n (n = 1, 2; M = Ge, Sn, Pb; M′=Cr, W). Eur. J. Inorg. Chem. 2000b, 4, 693–702.Suche in Google Scholar
Atwood, D. A.; Jegier, J. A.; Martin, K. J.; Rutherford, D. Tetradentate-N2 O2 ligand complexes of Tin(II). X-ray crystal structure of [N,N′-(1,2-ethylene)bis(salicylaldamine)]tin(II), (SaleanH2 Sn). J. Organomet. Chem. 1995, 503, C4–C7.Suche in Google Scholar
Berends, T.; Iovkova, L.; Bradtmöller, G.; Oppel, I.; Schürmann, M.; Jurkschat, K. LSn(OCH2 CH2)2 NR (L = lone pair, W(CO)5; R = Me, t-Bu). The molecular structures of 5-aza-2,8-dioxa-1-stannabicyclo[3.3.0]1.5 octanes and their tungstenpentacarbonyl complexes. Z. Anorg. Allg. Chem. 2009, 635, 369–374.Suche 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.Suche in Google Scholar
Calligaris, M.; Nardin, G.; Randaccio, L. Crystal and molecular structure of [NN′-ethylenebis(-salicylideneiminato)]dimethyltin(IV). J. Chem. Soc., Dalton Trans. 1972, 18, 2003–2006.10.1039/DT9720002003Suche in Google Scholar
ConQuest 1.16 (released December 2013) Search of the Cambridge crystal structure database (CSD).Suche in Google Scholar
Darensbourg, D.J.; Ganguly, P.; Billodeaux, D. Ring-opening polymerization of trimethylene carbonate using aluminum(III) and tin(IV) salen chloride catalysts. Macromolecules 2005, 38, 5406–5410.Suche in Google Scholar
Dey, D. K.; Das, M. K.; Noth H. Synthesis, spectroscopy and structure of diorganotin(IV) Schiff base complexes. Z. Naturforsch., B: Chem. Sci. 1999a, 54, 145–154.Suche in Google Scholar
Dey, D. K.; Saha, M. K.; Das, M. K.; Bhartiya, N.; Bansal, R. K.; Rosair, G.; Mitra, S. Synthesis and characterization of diorganotin(IV) complexes of tetradentate Schiff bases: crystal structure of n-Bu2Sn(Vanophen). Polyhedron 1999b, 18, 2687–2696.10.1016/S0277-5387(99)00171-0Suche 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.Suche in Google Scholar
Fjeldberg, T.; Hope, H.; Lappert, M. F.; Power, P. P.; Thorne, A. J. Molecular structures of the main group 4 metal(II)-bis(trimethylsilyl)-amides M[N(SiMe3)2]2 in the crystal (X-ray) and vapour (gas-phase electron diffraction). Chem. Commun. 1983, 639–641.10.1039/c39830000639Suche in Google Scholar
Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A. et al. Gaussian 09 , revision D.01, Gaussian, Inc.: Wallingford, CT, 2009.Suche in Google Scholar
González-García, G.; Álvarez, E.; Marcos-Fernández, Á.; Gutiérrez, J. A. Hexacoordinated oligosilanes from a hexacoordinated silicon(IV) complex containing an O,N,N,O salen-type and thiocyanato-N ligands. Inorg. Chem. 2009, 48, 4231–4238.Suche in Google Scholar
Iovkova-Berends, L.; Berends, T.; Dietz, C.; Bradtmöller, G.; Schollmeyer, D.; Jurkschat, K. Syntheses, structures and reactivity of new intramolecularly coordinated tin alkoxides based on an enantiopure ephedrine derivative. Eur. J. Inorg. Chem. 2011, 3632–3643.10.1002/ejic.201100352Suche in Google Scholar
Iovkova-Berends, L.; Berends, T.; Zöller, T.; Bradtmöller, G.; Herres-Pawlis, S.; Jurkschat, K. Tin(II) and tin(IV) compounds with scorpion-shaped ligands – intramolecular N→Sn vs. intermolecular O→Sn coordination. Eur. J. Inorg. Chem. 2012a, 3191–3199.Suche in Google Scholar
Iovkova-Berends, L.; Berends, T.; Zöller, T.; Schollmeyer, D.; Bradtmöller, G.; Jurkschat, K. Trapping molecular SnBr2(OH)2 by tin alkoxide coordination: syntheses and molecular structures of [MeN(CH2 CMe2 O)2 SnBr2]2·SnBr2(OH)2 and RN(CH2 CMe2 O)2 SnL [R = Me, n-octyl; L = lone pair, Cr(CO)5, W(CO)5, Fe(CO)4, Br2]. Eur. J. Inorg. Chem. 2012b, 3463–3473.Suche in Google Scholar
Jing, H; Edulji, S. K.; Gibbs, J. M.; Stern, C. L.; Zhou, H.; Nguyen, S. T. (Salen)tin complexes: syntheses, characterization, crystal structures, and catalytic activity in the formation of propylene carbonate from CO2 and propylene oxide. Inorg. Chem. 2004, 43, 4315–4327.Suche in Google Scholar
Kuchta, M. C.; Hahn, J. M.; Parkin, G. Divalent tin and lead complexes of a bulky salen ligand: the syntheses and structures of [salent-Bu,Me]Sn and [salent-Bu,Me]Pb. J. Chem. Soc., Dalton Trans. 1999, 3559–3563.10.1039/a905490aSuche in Google Scholar
Lippe, K.; Gerlach, D.; Kroke, E.; Wagler, J. Hypercoordinate organosilicon complexes of an ONN‘O’ chelating ligand: regio- and diastereoselectivity of rearrangement reactions in Si-salphen-systems. Organometallics 2009, 28, 621–629.Suche 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.Suche in Google Scholar
Mucha, F.; Böhme, U.; Roewer, G. Preparation, first X-ray structure analysis and reactivity of hexacoordinate silicon compounds with a tetradentate azomethine ligand. Chem. Commun. 1998, 12, 1289–1290.Suche in Google Scholar
Mucha, F.; Haberecht, J.; Böhme, U.; Roewer, G. Hexacoordinate silicon-azomethine complexes: synthesis, characterization, and properties. Monatsh. Chem. 1999, 130, 117–132.Suche 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.Suche in Google Scholar
Schenk, K. J.; Meghdadi, S.; Amirnasr, M.; Habibi, M. H.; Amiri, A.; Salehi, M.; Kashi, A. Co(III) complexes of Me-salpn and Me-salbn and the ring size effect on the coordination modes and electrochemical properties: The crystal structures of trans-[CoIII(Me-salpn)(py)2]PF6 and cis-a-[CoIII(Me-salbn)(4-Mepy)2]BPh4 · 4-Mepy. Polyhedron 2007, 26, 5448–5457.Suche in Google Scholar
Seiler, O.; Burschka, C.; Fischer, M.; Penke, M.; Tacke, R. Synthesis and structural characterization of novel neutral hexacoordinate silicon(IV) complexes with SiO2 N4 skeletons containing cyanato-N or thiocyanato-N Ligands. Inorg. Chem. 2005, 44, 2337–2346.Suche in Google Scholar
Sheldrick, G. M. SHELXL97. A computer program for crystal structure refinement. Version WinGX©, 1993–1997, Release 97-2. 1997a. University of Göttingen, Germany.Suche 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 of Göttingen, Germany.Suche in Google Scholar
Teoh, S.-G.; Yeap, G.-Y.; Loh, C.-C.; Foong, L.-W.; Teo, S.-B.; Fun, H.-K. Inner coordination sphere tin(IV) complexes with some O,N,N-terdentate{N-(2-hydroxybenzaldehyde)-1-amino-2-phenyleneimine and N-(2-hydroxy-1-naphthaldehyde)-1-amino-2-phenyleneimine} and O,N,N,O-quadridentate{N,N′-bis(2-hydroxybenzaldehyde)-1,2-phenylenediimine and N,N′-bis(2-hydroxy-1-naphthaldehyde)-1,2-phenylenediimine} Schiff bases. Polyhedron 1997, 16, 2213–2221.Suche in Google Scholar
van den Bergen, A. M.; Cashion, J. D.; Fallon, G. D.; West, B. O. Crystal structures, mößbauer spektra und reaktivity of Sn(II)-salicylideneimines. Aust. J. Chem. 1990, 43, 1559–1571.Suche in Google Scholar
Wagler, J.; Roewer, G. Intramolecular interligand charge transfer and a red hexacoordinate Si-complex with salen-type ligand vs. colorless tetracoordinate salen-Si-complexes with similar substituents. Inorg. Chim. Acta 2007, 360, 1717–1724.10.1016/j.ica.2006.09.006Suche in Google Scholar
Wagler, J.; Böhme, U.; Roewer, G. Silicon-enamine complexes: pentacoordinate silicon compounds. Angew. Chem. Int. Ed. 2002, 41, 1732–1734.Suche in Google Scholar
Wagler, J.; Doert, T.; Roewer, G. Synthesis of amines from imines in the coordination sphere of silicon—surprising photo-rearrangement of hexacoordinate organosilanes. Angew. Chem. Int. Ed. 2004, 43, 2441–2444.Suche in Google Scholar
Wagler, J.; Böhme, U.; Brendler, E.; Blaurock, S.; Roewer, G. Novel hexacoordinate diorganosilanes with salen-type ligands: molecular structure versus 29Si NMR chemical shifts. Z. Anorg. Allg. Chem. 2005a, 631, 2907–2913.Suche in Google Scholar
Wagler, J.; Böhme, U.; Brendler, E.; Thomas, B.; Goutal, S.; Mayr, H.; Kempf, B.; Remennikov, G. Ya.; Roewer, G. Switching between penta- and hexacoordination with salen-silicon-complexes. Inorg. Chim. Acta 2005b, 358, 4270–4286.10.1016/j.ica.2005.03.036Suche in Google Scholar
Wagler, J.; Roewer, G. First X-ray structures of ethylene bridged neutral dimeric hexacoordinate silicon complexes with tetradentate salen-type ligands. Z. Naturforsch. B 2005c, 60, 709–714.10.1515/znb-2005-0702Suche in Google Scholar
Wagler, J.; Gerlach, D.; Böhme, U.; Roewer, G. Intramolecular inter-ligand charge transfer in hexacoordinate silicon complexes. Organometallics 2006, 25, 2929–2933.Suche in Google Scholar
Wagler, J.; Roewer, G.; Gerlach, D. Photo-driven Si-C bond cleavage in hexacoordinate silicon complexes. Z. Anorg. Allg. Chem. 2009, 635, 1279–1287.Suche in Google Scholar
Wang, Q.; Ding, R.; Wang, N.; Zhang, D.; Li, J. Synthesis and characterization of diorganotin(IV) complexes bearing binary Schiff-bases and crystal structure of n-Bu2 SnL1. Z. Anorg. Allg. Chem. 2010, 636, 861–864.Suche in Google Scholar
Westerhausen, M.; Schneiderbauer, S; Kneifel, A. N.; Söltl, Y.; Mayer, P.; Nöth, H.; Zhong, Z; Dijkstra,P. J.; Feijen, J. Organocalcium compounds with catalytic activity for the ring-opening polymerization of lactones. Eur. J. Inorg. Chem. 2003, 18, 3432–3439.Suche in Google Scholar
Yearwood, B.; Parkin, S.; Atwood, D. A. Synthesis and characterization of organotin Schiff base chelates. Inorg. Chim. Acta 2002, 333, 124–131.Suche in Google Scholar
Zöller, T.; Iovkova-Berends, L.; Berends, T.; Dietz, C.; Bradtmöller, G.; Jurkschat, K. Intramolecular N→Sn coordination in tin(II) and tin(IV) compounds based on enantiopure ephedrine derivatives. Inorg. Chem. 2011, 50, 8645–8653.Suche in Google Scholar
Zöller, T.; Lutter, M.; Berends, T.; Jurkschat, K. The 2,8-dioxa-5-aza-1-sila-bicyclo[3.3.01.5]octane PhN(CH2 CH2 O)2 SiH2 as reducing reagent: synthesis and molecular structure of PhN(CH2 CH2 O)2 Sn. Main Group Met. Chem. 2013, 36, 77–82.Suche in Google Scholar
©2014 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.
Artikel in diesem Heft
- Frontmatter
- Research articles
- Molecular structures of Sn(II) and Sn(IV) compounds with di-, tri- and tetramethylene bridged salen* type ligands
- Synthesis and structures of Cu-Cl-M adducts (M=Zn, Sn, Sb)
- Synthetic strategy for the incorporation of Bu2Sn(IV) into fluorinated β-diketones/benzoylacetone and sterically demanding heterocyclic β-diketones and spectroscopic characterization of hexacoordinated complexes of Bu2Sn(IV)
- Reactivity of bis(cyclohexylammonium) 4-nitrophenylphosphate with SnMe3 Cl. X-ray structure of 4-NO2 C6 H4 PO4(SnMe3)2·H2 O
- Structural elucidation of novel mixed ligand complexes of 2-thiophene carboxylic acid [M(TCA)2(H2O)x(im)2] [x=2 M: Mn(II), Co(II) or Cd(II), x=0 Cu(II)]
- Short Communication
- Synthesis and structure of the first tin(II) amidinato-guanidinate [DippNC(nBu)NDipp]Sn{pTol-NC[N(SiMe3)2]N-pTol}
Artikel in diesem Heft
- Frontmatter
- Research articles
- Molecular structures of Sn(II) and Sn(IV) compounds with di-, tri- and tetramethylene bridged salen* type ligands
- Synthesis and structures of Cu-Cl-M adducts (M=Zn, Sn, Sb)
- Synthetic strategy for the incorporation of Bu2Sn(IV) into fluorinated β-diketones/benzoylacetone and sterically demanding heterocyclic β-diketones and spectroscopic characterization of hexacoordinated complexes of Bu2Sn(IV)
- Reactivity of bis(cyclohexylammonium) 4-nitrophenylphosphate with SnMe3 Cl. X-ray structure of 4-NO2 C6 H4 PO4(SnMe3)2·H2 O
- Structural elucidation of novel mixed ligand complexes of 2-thiophene carboxylic acid [M(TCA)2(H2O)x(im)2] [x=2 M: Mn(II), Co(II) or Cd(II), x=0 Cu(II)]
- Short Communication
- Synthesis and structure of the first tin(II) amidinato-guanidinate [DippNC(nBu)NDipp]Sn{pTol-NC[N(SiMe3)2]N-pTol}