Abstract
Reaction of the chelating imino-pyridine ligand SIMPY, (SIMPY=2-(DippN=CH)-C5H4N), Dipp=2,6-iPr2-C6H3, with germanium(II) and tin(II) halides provides the respective neutral complexes [SIMPY·EX2] (EX2: E=Ge, X=Cl, Br; E=Sn, X=Cl, Br, I). The method is readily extendable to give the tin(II) triflate complex [SIMPY·Sn(OTf)2] (OTf, triflate=CF3SO3−). In the solid state, the neutral compounds [SIMPY·EX2] exist as monomers, in which the four-coordinate tetrel atoms feature a slightly distorted disphenoidal geometry around germanium and tin. Reaction of the tridentate imino-pyridine ligand DIMPY, (DIMPY=2,6-(DippN=CH)2-C5H3N) with Sn(OTf)2 provided access to a neutral tin(II) complex. Similar to the previously reported reactions leading to the germanium and tin chloride complexes [DIMPY·SnCl]+[SnCl3]−, and [Me2DIMPY·EX]+[EX3]− (Me2DIMPY=2,6-(DippN=C(Me))2-C5H3N, E=Ge, Sn; X=Cl), the reactions of DIMPY with GeX2·dioxane (X=Cl, Br) and SnX2 (X=Br, I) yielded Ge(II) and Sn(II) based ion pairs [DIMPY·EX]+[EX3]− (E=Ge, X=Cl, Br; E=Sn, X=Br, I) as a consequence of spontaneous dissociation of the group 14 dihalides. The tetrel atoms in the cationic parts in [DIMPY·EX]+[EX3]− are four-coordinate as one halide substituent is replaced by the coordination of a second imino donor group from the ligand. The anionic fragments adopt a pyramidally, tri-coordinate geometry. In contrast, the DIMPY tin(II) ditriflate complex crystallizes with two independent, neutral molecules per asymmetric unit, in which one of the tin centers is five- coordinate by interaction with three donor sites of the chelating bis(imino)pyridine ligand and two additional contacts towards the oxygen atoms of the triflate counter-anions. In the second crystallographically independent complex the tin atom is six-coordinate with a slightly distorted octahedral geometry via interaction with THF as an additional donor molecule. All compounds reported were studied by means of multinuclear NMR spectroscopy. In addition, the solid state structures of the complexes [SIMPY·EX2] (EX2: E=Ge, X=Cl, Br; E=Sn, X=Cl, Br, I), the ion pairs [DIMPY·EX]+[EX3]− (E=Ge, X=Cl; E=Sn, X=Br) and the tin(II) ditriflate [DIMPY·Sn(OTf)2] were authenticated by means of single-crystal X-ray diffraction analyses. Moreover, [DIMPY·Sn(OTf)2] was investigated by 119Sn Mössbauer spectroscopy.
Dedicated to: Professor Dietrich Gudat on the occasion of his 60th birthday.
1 Introduction
Redox non-innocent pyridine ligands with chelating, imino-functionalized side arms in ortho-positions have found widespread application in the coordination chemistry of low oxidation state transition metals, for example in 2,6-diimino functionalized-pyridine complexes of iron [1], [2], [3], [4], [5] or cobalt [6], but also in the first row transition metal complexes of the monosubstituted SIMPY ligand 2-(DippN=C(Me))-C5H3N [7].
To a smaller extent, these pincer ligands were employed in the synthesis and stabilization of main group element compounds as they can act as two- or three-fold chelating ligands and hence stabilize Lewis acidic s- and p-block metal centers. Examples of main group element compounds include derivatives of group 1 [8], [9] and 2 [10], [11], [12], group 13 [13], [14], [15], [16], [17], [18], [19], [20], [21], [22], [23], [24], [25], [26], [27], group 14 [28], and group 15 [29], [30], [31]. As a direct consequence of their uncharged nature, iminopyridine ligands are interesting candidates to stabilize cationic low coordinate, main group element complexes. Since the seminal discovery of crown ether stabilized germanium dications [32], these complexes are not only of interest due to their unusual coordination behavior but stirred interest in the context of potential small molecule activation. Cationic iminopyridine compounds of groups 15 [33] and 16 elements include an arsenic(I) cation [34] and dicationic group 16 (S2+, Se2+, Te2+) [35] complexes. Examples of iminopyridine-based heavier group 14 element compounds, in which the tetrel centers are in a formal oxidation state of +II include the ion pairs [Me2DIMPY·EX]+[EX3]− (Me2DIMPY=2,6-(DippN=C(Me))2-C5H3N, E=Ge; Sn, X=Cl) [36], [Ph2DIMPY·EX]+[EX3]− (Ph2DIMPY=2,6-(ArN=C(Ph))2-C5H3N, Ar=2,5-tBu2C6H3, 2,6-Me2C6H3, E=Sn, X=Cl, Br) [37] and [6-MeOSIMPY·EX]+[EX3]− (6-MeOSIMPY=2-(DippN=C(Me))-6-MeO-C5H3N), E=Ge, Sn; X=Cl) [38] and furthermore cationic tin complexes of metallocene functionalized DIMPY ligands [39]. Recently, it has been shown that the SnCl3− ions in these compounds can act as Lewis bases and are able to complex transition metal fragments such as the group 6 pentacarbonyl M(CO)5 (M=Cr, Mo, W) or [η6-arene RuCl]+ units (arene=C6H6, iPr-C6H4-Me) [40]. Moreover, complexes of Me2DIMPY with tin(II) triflate have been obtained [41]. Recently, we reported on the ion pair [DIMPY·SnCl]+[SnCl3]− and the unusual complex [DIMPY·Sn], in which the tin center exists in the formal oxidation state zero [42] (Fig. 1).

Germanium and tin mono- and diiminopyridine complexes.
Although formation of low oxidation state group 14 cationic complexes [L·EX]+ from EX2 precursors is well documented in literature, these ion pair complexes are rather exceptional and attracted considerable attention as typically uncharged complexes [L·EX2] are obtained from group 14 dihalide precursors. Therefore, it is even more surprising that no neutral 1:1 complexes [L·EX2] based on chelating iminopyridine ligands have been reported yet. Herein, we exemplify the formation of such neutral compounds [SIMPY·EX2] together with additional examples of ion pairs [DIMPY·EX]+[EX3]− and show that a trend towards dissociation is already apparent in the neutral SIMPY species. The presence of an additional donor group in the ligand induces dissociation of the group 14 dihalides. The tin ditriflate species [DIMPY·SnOTf2] is remarkable in this context as no auto-ionization but the formation of a neutral 1:1 complex is observed. In this context it is noteworthy that the recently reported tin ditriflates [Me2DIMPY·SnOTf2] and [MeOSIMPY·SnOTf2] also do not show auto-ionization in the solid state. The mixed species [MeOSIMPY·SnCl]+OTf−, however, features a weakly coordinated triflate group with a very long Sn-OTf distance in the solid state [41].
2 Results and discussion
2.1 Synthesis
Addition of GeX2·dioxane (X=Cl, Br) and SnX2 (X=Cl, Br, I, OTf; OTf=triflate, CF3SO3−) to solutions of one equivalent of the chelating imino pyridine ligand SIMPY in THF results in an immediate color change from colorless to yellow or deep orange red depending on the nature of the group 14 element and of the anion. Heavier homologues cause the expected bathochromic shift. Removal of the solvent gave the neutral SIMPY·EX2 complexes 1–6 (1: E=Ge, X=Cl, 2: E=Ge, X=Br, 3: E=Sn, X=Cl, 4: E=Sn, X=Br, 5: E=Sn, X=I, 6: E=Sn, X=OTf) in essentially quantitative yields as pure products judged by multinuclear NMR spectroscopy (Scheme 1). Recrystallization from THF or THF-benzene mixture provided X-ray quality crystals of the dihalide complexes 1–5. Despite numerous attempts, for the ditriflate compound 6 no crystals of sufficient quality for single crystal X-ray diffraction analysis were obtained (Scheme 1).

Syntheses of the neutral complexes 1–6, SIMPY EX2.
Similar to the formation of previously reported 12, [DIMPY·SnCl]+[SnCl3]−, reaction of the DIMPY ligand with GeX2·dioxane (X=Cl, Br) and SnX2 (X=Br, I) in THF following a 1:2 stoichiometry results in the precipitation of the respective ion pairs [DIMPY·EX]+[EX3]−, 7–10, (7: E=Ge, X=Cl, 8: E=Ge, X=Br, 9: E=Sn, X=Br, 10: E=Sn, X=I). In contrast, reaction of the DIMPY ligand with tin(II) ditriflate, Sn(OTf)2, yields a product with 1:1 stoichiometry 11 as established by single crystal X-ray diffraction analysis (Scheme 2). Investigations to react DIMPY with one equivalent of SnX2 lead to the formation of 0.5 equivalents of complexes 7–10 and recovery of 0.5 eq. of the uncoordinated ligand.
![Scheme 2: Formation of the DIMPY complexes [DIMPY EX]+[EX3]− and [DIMPY Sn(OTf)2].](/document/doi/10.1515/znb-2017-0128/asset/graphic/j_znb-2017-0128_scheme_002.jpg)
Formation of the DIMPY complexes [DIMPY EX]+[EX3]− and [DIMPY Sn(OTf)2].
The ion pairs 7–10 and 12 are significantly less soluble in ethereal solvents and chlorohydrocarbons such as dichloromethane or chloroform in comparison to the neutral complexes 1–6 and 11. It is noteworthy that attempted syntheses employing more than one equivalent of EX2 for compounds 1–6 and 11 did not result in the formation of ion pairs, but resulted in the recovery of unreacted group 14 dihalide salts. Vice versa, attempts with stoichiometries different from 1:2 in the synthesis of the ion pairs 7–10 and 12 did not produce neutral, undissociated complexes, but left a respective amount of unreacted DIMPY ligand in solution. The only exception for the formation of ion pairs with the DIMPY ligand is the respective neutral tin ditriflate complex 11. Similarly, the ortho-methoxy substituted ligand 6-MeOSIMPY (6-MeOSIMPY=2-(ArN=C(Me))-6-MeO-C5H3N)) induces ion pair formation as in the complexes [MeOSIMPY·ECl]+[ECl3]−, E=Ge, Sn, although the distances between the germanium and tin atoms and the methoxy sidearm are long and exceed the sum of the van der Waals radii of oxygen and germanium or tin, respectively [38].
2.2 Spectroscopy and characterization
All compounds have been characterized by 1H, 13C and in the case the tin complexes, by 119Sn NMR spectroscopy in solution and via elemental analyses. The solid state structures of compounds 1–5, 7, 9 and 11 were authenticated by single crystal X-ray diffraction analyses. 11 was additionally investigated by 119Sn Mössbauer spectroscopy.
2.2.1 NMR spectroscopy
The 119Sn NMR spectra of the tin complexes 3–5, 9, 10 and 12 clearly demonstrate the dependence of the 119Sn NMR shifts on the nature of the halides, which corroborates the absence of complete dissociation of the Sn–halide bonds upon dissolving the samples. The 119Sn signals of the SIMPY complexes 3–5 move to lower field upon descending group 17. [SIMPY·SnCl2], 3, shows a resonance at −216 ppm in THF, the resonance of the respective bromide 4 is shifted to −80 ppm and the iodide resonates at +317 ppm. The shifts of 3 and 4 correlate well with literature data for tin dichloride (−236 ppm in THF) and tin dibromide (−72 ppm in dimethoxyethane) dissolved in strong donor solvents.
The tin-based ion pairs [DIMPY·SnX]+[SnX3]−9 (X=Br), and 12 (X=Cl) exhibit two pairs of well distinguishable 119Sn NMR resonances in solution each, at +98 ppm [DIMPY·SnBr]+/−406 ppm [SnBr3]− and −452 [DIMPY·SnCl]+/−58 ppm [SnCl3]−. The shifts of the chloride derivative 12 correlate well with those observed for the previously reported compounds [Me2DIMPY SnCl]+[SnCl3]− (−435/−73 ppm) [36] and [6-MeOSIMPY SnCl]+[SnCl3]− (−330/−73 ppm) [38] and related transition metal complexes reported by Jambor, Jurkschat and coworkers [40]. In comparison to the chlorides, the positions of the resonances of the bromides switch places. The signal for the cationic fraction is shifted downfield and is observed at −98 ppm, the [SnBr3]− ion resonates at −406 ppm, which correlates reasonably well with literature data [36]. Despite numerous attempts including variation of the concentration, solvent and sample temperature, no 119Sn NMR resonances for the iodide 10 were observed. Both tin(II) ditriflates 6 and 11 show only one 119Sn NMR signal in solution at −27 ppm (6) and −46 ppm (11). These values agree reasonably well with data for [6-MeOSIMPYSnOTf2] (δ=−78 ppm) [41]. Together with the significant line broadening effects observed in 1H spectra and rather long Sn–O distances in the solid state, NMR spectroscopic evidence points towards rapid dissociation/association processes in solution. Although low-temperature NMR experiments were conducted, we were unable to resolve the exchange processes on the NMR timescale. As shown in our previous publication, the 1H NMR shift of the imino proton and the resonances of the pyridine fragment are diagnostic for the formation of the group 14 element complexes [42]. In all complexes 1–6, the 1H NMR shifts of these signals differ notably from those of the free ligand and therefore provide clear evidence for the existence of the complex species in solution. Dissociation of the complexes to yield free ligand and solvates of the tetrel dihalides can be ruled out even when strong donor solvents (e.g. THF, dimethoxyethane) are used in sample preparation. In all SIMPY complexes, the 1H resonances of the imino proton ArN=CH is shifted downfield with respect to the free ligand (8.31 ppm). The most pronounced effect is observed for the tin ditriflate (9.43 ppm) and the germanium and tin dichlorides (8.65 ppm in 1 and 8.50 ppm in 3). The bromides 2 and 4 and the iodide 5 display less significant shift changes. Similar trends are observed in the DIMPY complexes 7–12, where once again the shift changes between free ligand and group 14 complexes decrease from chloride to bromide and iodide. The tin ditriflates, however, are different in this context as the SIMPY complex shows the most pronounced downfield shift, whereas the DIMPY complex exhibits a slightly upfield shifted 1H resonance at 8.18 ppm for the ArN=CH proton.
The methyl resonances of the iso-propyl groups in the SIMPY complexes 1–5 are broad and do not show two distinct resonances in 1H and 13C spectra as it is usually observed in related Dipp-substituted complexes. The ditriflate 6, however, shows two distinct 13C resonances for these methyl groups, which are nevertheless isochronous in the 1H spectrum. Similarly, only the methyl resonances in chloride complexes 7 and 12 and in the tin bromide 9 appear as two sets of signals in the 1H and 13C spectra, but are not resolved in the germanium bromide 8 and the tin iodide 10. Spectroscopic evidence therefore points towards a rapid exchange of the axial and equatorial halide substituents in the [SIMPY·EX2] complexes and a similar positional exchange in the cationic fragments [DIMPY·EX]+, which is faster for the heavier halogen atoms.
2.2.2 X-ray crystallography
The solid-state structures of the complexes 1–5, 7, 9 and 11 were authenticated by means of single crystal X-ray diffraction analyses. All SIMPY based compounds crystallize as undissociated, neutral, binary complexes [L·EX2]. In contrast, the DIMPY complexes of the group 14 dihalides form separated ion pairs [DIMPY·EX]+[EX3]− upon auto-ionization. The DIMPY tin ditriflate 11, however, crystallizes as a neutral, undissociated complex with two independent molecules per asymmetric unit in which one of the tin centers is five-coordinate, while the second tin atom is six-coordinate via interaction with a THF donor molecule. A summary of crystallographic data is given in Table 1.
Crystallographic data for compounds 1–5, 7, 9, 11.
Compound | 1 [SIMPYGeCl2] | 2 [SIMPYGeBr2] | 3 [SIMPYSnCl2] | 4 [SIMPYSnBr2] | 5 [SIMPYSnI2] | 7 [DIMPYGeCl][GeCl3] | 9 [DIMPYSnBr][SnBr3]·4THF | 11 [DIMPYSnOTf2]2·3THF |
---|---|---|---|---|---|---|---|---|
CCDC number | 1557346 | 1557344 | 1557339 | 1557341 | 1557345 | 1557340 | 1557342 | 1557343 |
Empirical formula | C18H22Cl2GeN2 | C18H22Br2GeN2 | C18H22Cl2SnN2 | C18H22Br2SnN2 | C18H22I2SnN2 | C31H39Cl4Ge2N3 | C47H69Br4N3O4Sn2 | C74H94F12N6O14S4Sn2 |
Formula weight | 409.87 | 498.79 | 455.97 | 544.89 | 638.87 | 740.63 | 648.54 | 1885.17 |
Temperature, K | 100(2) | |||||||
Wavelength λ, Å | 0.71073 | |||||||
Crystal system | Monoclinic | Monoclinic | Monoclinic | Monoclinic | Triclinic | Monoclinic | Monoclinic | Monoclinic |
Space group | P21/n | P21/n | P21/n | P21/n | P1̅ | P21/c | P21/c | P21/n |
a, Å | 9.4954(5) | 9.5142(6) | 9.5486(4) | 9.5680(7) | 8.035(13) | 17.0347(9) | 19.8018(18) | 17.9231(9) |
b, Å | 13.6124(8) | 13.8330(8) | 13.9496(8) | 14.1891(10) | 10.531(13) | 13.1207(6) | 10.6387(9) | 15.3185(7) |
c, Å | 14.6014(8) | 14.6809(9 | 14.3876(7) | 14.5150(11) | 13.86(2) | 17.5295(8) | 26.397(2) | 30.4011(13) |
α, deg | 90 | 90 | 90 | 90 | 102.36(5) | 90 | 90 | 90 |
β, deg | 94.634(2) | 93.311(2) | 92.019(2) | 92.079(3) | 94.72(4) | 118.813(2) | 106.747(4) | 94.132(2) |
γ, deg | 90 | 90 | 90 | 90 | 107.43(5) | 90 | 90 | 90 |
V, Å3 | 1881.14(18) | 1928.9(2) | 1915.23(16) | 1969.3(2) | 1079(3) | 3432.9(3) | 5325.1(8) | 8325.1(7) |
Z | 4 | 4 | 4 | 4 | 2 | 4 | 8 | 4 |
ρcalcd, g cm−3 | 1.45 | 1.72 | 1.58 | 1.84 | 1.97 | 1.43 | 1.62 | 1.50 |
μ, mm−1 | 1.9 | 5.7 | 1.6 | 5.4 | 4.0 | 2.1 | 4.0 | 0.8 |
F(000), e | 840 | 984 | 912 | 1056 | 600 | 1512 | 2576 | 3856 |
Crystal size, mm3 | 0.22×0.17×0.12 | 0.18×0.13×0.09 | 0.13×0.12×0.11 | 0.36×0.24×0.15 | 0.25×0.18×0.07 | 0.23×0.19×0.16 | 0.84×0.48×0.17 | 0.61×0.38×0.17 |
θ range, deg | 2.05–30.00 | 2.49–28.89 | 2.03–28.99 | 2.01–30.00 | 2.69–27.00 | 1.36–27.00 | 1.66–28.00 | 1.28–27.00 |
Index ranges h,k,l | −13≤h≤14 | −12≤h≤12 | −12≤h≤12 | −13≤h≤13 | −10≤h≤10 | −21≤h≤21 | −26≤h≤26 | −22≤h≤22 |
−20≤k≤20 | −18≤k≤18 | −18≤k≤19 | −19≤k≤19 | −13≤k≤13 | −16≤k≤16 | −14≤k≤14 | −19≤k≤19 | |
−22≤l≤22 | −19≤l≤19 | −14≤l≤19 | −20≤l≤20 | −17≤l≤17 | −20≤l≤22 | −34≤l≤33 | −38≤l≤34 | |
Reflexions collected | 75 969 | 43 853 | 24 438 | 41 834 | 24 416 | 1 27 366 | 1 49 107 | 1 47 268 |
Refl. unique/Rint | 6975/0.0344 | 5071/0.0370 | 5068/0.0285 | 5739/0.0675 | 4685/0.0437 | 7571/0.0677 | 12 851/0.0979 | 18 163/0.0752 |
Completeness to θmax, % | 99.9 | 99.9 | 99.7 | 99.9 | 99.2 | 99.6 | 100.0 | 99.9 |
Absorption correction | Multi-scan/Sadabs | |||||||
Refinement method | Full-matrix least-squares on F2 | |||||||
Data/parameters | 6975/212 | 5071/212 | 5068/212 | 5739/212 | 4685/212 | 7571/382 | 12 851/549 | 18 163/1090 |
R1/wR2 [I>2σ(I)] | 0.0216/0.0557 | 0.0194/0.0455 | 0.0304/0.0728 | 0.0326/0.0822 | 0.0252/0.0444 | 0.0643/0.1352 | 0.0624/0.1726 | 0.0714/0.1795 |
R1/wR2 (all data) | 0.0280/0.0589 | 0.0252/0.0478 | 0.0358/0.0761 | 0.0343/0.0834 | 0.0485/0.0507 | 0.0840/0.1455 | 0.0723/0.1789 | 0.1022/0.1957 |
GoF | 1.029 | 1.038 | 1.071 | 1.101 | 1.015 | 1.400 | 1.053 | 1.115 |
Δρfin (max/min), e Å−3 | 0.53/−0.32 | 0.67/−0.35 | 2.35/−1.08 | 1.44/−1.40 | 0.55/−0.46 | 0.72/−0.74 | 1.41/−2.03 | 2.75/−0.96 |
The complexes [SIMPY·EX2] (EX2=GeCl2, GeBr2, SnCl2, SnBr2) crystallize in the monoclinic space group P21/n whereas the iodide SIMPY SnI2 crystallizes in the triclinic space group P1̅. All tetrel centers in the SIMPY complexes are four-coordinate and adopt a distorted disphenoidal structural motif (AB4Ë-type structure according to VSEPR rules) around the tetrel atom, which is coordinated by two halide ions and two nitrogen donors of the SIMPY ligand. Both nitrogen atoms and one halide ion are located in plane with the central tetrel atom while the second halide ion is located at right angle to this plane (Fig. 2).

Schematic drawing of the group 14 element coordination in the complexes 1–5, 7 and 9 with the numbering scheme used in the text.
In all cases, two distinctly different E–N bond lengths are observed, where the E–N2 separation between the group 14 atom and the nitrogen atom of the pyridine ring is significantly shorter than the E–N1 distance between the tetrel and the imino nitrogen of the side arm. Furthermore, the tetrel halogen distances E–X1 (perpendicular to the N1/N2/E plane) and E–X2 (in plane) differ notably in all SIMPY group 14 derivatives.
In all compounds studied crystallographically, the longer E–X2 bond is found in trans position with respect to the weaker imino nitrogen donor atom N1 and lies in the plane E–N1–C1–C2–N2. The shorter E–X1 bond invariably involves the second halide atom, which is located in an almost perpendicular position with respect to the same plane. A summary of bond lengths and angles is given in Table 2.
Selected bond lengths (Å) and angles (deg) for compounds 1–5.
1 [SIMPY·GeCl2] | 2 [SIMPY·GeBr2] | 3 [SIMPY·SnCl2] | 4 [SIMPY·SnBr2] | 5 [SIMPY·SnI2] | |
---|---|---|---|---|---|
E1–N1 | 2.4900(10) | 2.437(1) | 2.565(2) | 2.522(1) | 2.517(6) |
E1–N2 | 2.1752(8) | 2.170(2) | 2.364(4) | 2.376(2) | 2.342(5) |
Δ(E–N1/2) | 0.315 | 0.267 | 0.292 | 0.176 | 0.175 |
E1–X2 | 2.4039(2) | 2.6104(4) | 2.5640(6) | 2.5955(3) | 2.902(1) |
E1–X1 | 2.2722(3) | 2.4323(4) | 2.4411(6) | 2.7478(3) | 2.997(1) |
Δ(E–X1/2) | 0.132 | 0.178 | 0.123 | 0.152 | 0.095 |
N1–C1 | 1.271(1) | 1.272(3) | 1.273(2) | 1.273(2) | 1.319(9) |
N1–E–N2 | 71.21(5) | 72.02(4) | 67.31(5) | 67.58(5) | 67.80(2) |
N1–E–X1 | 162.46(4) | 164.83(3) | 153.83(4) | 155.97(3) | 158.50(10) |
X1–E–X2 | 91.29(1) | 91.04(1) | 88.43(2) | 88.25(1) | 92.59(2) |
N1–E–X2 | 83.23(3) | 84.43(5) | 82.14(4) | 82.96(3) | 84.70(10) |
N2–E–X1 | 91.32(2) | 91.15(5) | 88.44(4) | 90.13(4) | 91.40(10) |
N2–E–X2 | 92.36(2) | 93.66(5) | 90.04(4) | 89.67(4) | 90.90(10) |
The largest difference in E–X bond lengths in compounds 1–5 is observed in the bromides 2 and 4. In 2, the Ge1–Br2 distance amounts to 2.6104(4) Å and is thus 0.1781 Å longer that the out-of-plane Ge1–Br1 bond. Somewhat less pronounced is the difference in 4, where 2.7478(2) Å for Sn1–Br2 and 2.5955(3) Å for Sn1–Br1 are found. Similarly, the two E–N distances vary notably, with a short E–N(pyridine) distance and an elongated E–N(imine) contact. A comparison of the chlorides 1 (Fig. 3) and 3 with the bromides 2 and 4 shows that shorter and therefore stronger E–N1 interactions as in 2 vs. 1 correlate to longer and weaker E–X2 distances in trans-position which ultimately results in the dissociation of the E–X2 bond in the DIMPY complexes and the modified SIMPY ligand based complexes [6-MeOSIMPY·EX]+[EX3]− (E=Ge; Sn, X=Cl) and hence formation of ion pairs [39]. The coordination in the tin diiodide 5, where the tin-halogen interaction is characterized as a soft-soft Lewis acid-base interaction, is different. Notably, it exhibits not only the shortest Sn1–N1 and Sn1–N2 bonds, but also the most symmetrical bonding situation with differences of 0.175 Å for the Sn–N bonds and only 0.095 Å for the two tin iodide bonds. The N1–C1 distance, which is diagnostic for the electron density transfer from the metal center towards the ligand does not vary significantly in the chlorides and bromides 1–4 and is similar to that of the uncomplexed ligand. In contrast, the C1–N1 distance is ca. 0.04 Å longer in the iodide complex 5.
![Fig. 3: Molecular structure of [SIMPY·GeCl2], 1, in the solid state. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are omitted for clarity.](/document/doi/10.1515/znb-2017-0128/asset/graphic/j_znb-2017-0128_fig_003.jpg)
Molecular structure of [SIMPY·GeCl2], 1, in the solid state. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are omitted for clarity.
With respect to the germanium dichloride complex 1, the germanium-nitrogen distances of 2.070(4) Å for Ge1–N1 and 2.045(4) Å for Ge1–N2 in the crystallographically characterized, methoxy substituted SIMPY complex [6-MeOSIMPY·GeCl]+[GeCl3]− are found to be short and more symmetrical [39]. Moreover, the Ge–Cl distance in the tricoordinate cation is also ca. 0.06 Å shorter than in 1. The differences are a direct consequence of the lower coordination number and the positive charge in [6-MeOSIMPY·GeCl]+ compared to 1.
The introduction of a second donor side arm as present in the DIMPY ligand apparently tips the balance towards auto-ionization and results in the formation of the ion pairs 7–10 and 12. Similar to [Me2DIMPY·EX]+[EX3]− [36] and [6-MeOSIMPY·EX]+[EX3]− [39], the bonding situation is characterized by rather short and symmetrical bonds E–N1 and E–N2 towards the imino nitrogen donors and an even shorter E–N3 distance. In the solid state, the halogen atom is invariably located below the N1–N2–N3–E plane and forms three angles N–E–X near 90°.
The molecular structures of 7, 9, previously reported 12 and the Me-DIMPY species [Me2DIMPY·EX]+[EX3]− [36] contain well separated ion pairs in which the anion [EX3]− is strictly pyramidal with a sum of angles of 286.8(2)° in 7 and 284.2(1)° in 9 (Fig. 4). The cations contain four-coordinate tetrel centers, with two longer E–N separations towards the imino nitrogen atoms N1 and N2 and a short E–N3 distance towards the pyridine nitrogen atom. The fourth coordination site is occupied by a halogen atom. In 7, the Ge1–N1 and Ge1–N2 distances are 2.255(6) Å and 2.316(6) A, whereas the Ge1–N3 distance is only 2.075(4) Å. In the tin bromide complex 9, the respective distances are 2.453(2) and 2.457(2) Å (Sn1–N1 and Sn1–N2) which are slightly elongated compared to the tin chloride complex 12 based on the identical ligand, where Sn1–N1 and Sn1–N2 are 2.414(2) and 2.453(2) Å. The Sn1–N3 distance in 9 is conversely slightly shorter at 2.280(3) Å compared to the 2.302(2) Å observed in 12. In any case, both the E–N(pyridine) and the E–N(imine) bonds observed in the DIMPY complexes are longer than in the SIMPY complexes. The N3–E–N1/N2 angles are more acute for the germanium complexes than for the respective tin complexes, which is a direct consequence of the smaller atomic radius of germanium.
![Fig. 4: Molecular structure of [DIMPY·SnBr]+[SnBr]−, 9, in the solid state. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are omitted for clarity.](/document/doi/10.1515/znb-2017-0128/asset/graphic/j_znb-2017-0128_fig_004.jpg)
Molecular structure of [DIMPY·SnBr]+[SnBr]−, 9, in the solid state. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are omitted for clarity.
Interestingly, the molecular structure of the tin ditriflate derivative 11 (space group P1̅, Fig. 5) is strikingly different, as is does not consist of ion pairs but of two structurally independent [DIMPY·Sn(OTf)2] complexes per asymmetric unit. One of the complexes features a five-coordinate tin(II) center with essentially symmetrical interactions between the tin atom and the imino nitrogen atoms (Sn2–N4: 2.515(5) and Sn2–N6 2.501(5) Å) and a rather short Sn–N5(pyridine) distance of 2.314(5) Å. Moreover, the triflate groups are both engaged in short, direct interactions with the tin center with distances of 2.379(8) and 2.394(7) Å and are located above and below the basal plane of the DIMPY ligand containing the tin atom.
The second, structurally independent molecule features a six-coordinate tin(II) center in which all substituents are located at large distances. The Sn–imino distances are notably different and amount to 2.477(6) and 2.772(6) Å. The Sn1–N3 distance between to the pyridine nitrogen atom is also longer with 2.357(6) Å. Furthermore the Sn–O distances towards the triflate groups differ in lengths, one being shorter (2.248(8) Å) than both Sn–O distances in the five-coordinate species while the other distance is significantly elongated (2.611(6) Å). The THF molecule is weakly coordinated as the Sn–O(THF) distance is 2.725 Å. The Sn–O distances are in good agreement with the distances found in [Me2DIMPY·SnOTf2] (2.350(2) and 2.554(2) Å) and in [MeOSIMPY·SnOTf2] (2.3822(9) and 2.443(2) Å). For comparison, the Sn–O distance in the mixed species [MeOSIMPY·SnCl]+OTf− is significantly elongated at 3.042(2) Å [41].
A summary of bond lengths and angles in the DIMPY complexes is given in Table 3.
Bond lengths (Å) and angles (deg) in 7, 9 and 11.
7 [DIMPY·GeCl]+[GeCl3]– | 9 [DIMPY·SnBr]+[SnBr3]– | 11 [DIMPY·SnOTf2] | ||
---|---|---|---|---|
5 coord. | 6 coord. | |||
E–X1 | 2.223(2) | 2.6013(7) | 2.379(8)/2.394(7) | 2.248(8)/2.611(8) |
E–N1 | 2.255(6) | 2.453(4) | 2.515(5) | 2.477(6) |
E–N2 | 2.316(6) | 2.457(4) | 2.501(5) | 2.772(6) |
E–N3 | 2.075(4) | 2.280(3) | 2.314(5) | 2.357(6) |
C1–N1 | 1.271(7) | 1.265(3) | 1.260(6) | 1.270(5) |
C2–N2 | 1.276(6) | 1.270(5) | 1.278(6) | 1.260(6) |
C1–C3 | 1.472(9) | 1.466(7) | 1.485(6) | 1.470(5) |
C2–C7 | 1.392(7) | 1.464(7) | 1.468(6) | 1.472(5) |
N1–E–N2 | 146.45(7) | 136.6(1) | 137.65(8) | 133.77(7) |
N1–E–X1 | 86.24(7) | 86.39(9) | 78.79(9)/80.22(9) | 74.44(8)/78.69(8) |
N2–E–X1 | 94.22(6) | 85.98(9) | 89.40(8)/91.03(9) | 92.71(9)/95.39(9) |
N3–E–X1 | 89.22(7) | 88.37(9) | 77.81(9)/74.38(8) | 79.95(9)/77.64(8) |
Σ(E′X3) in [EX3]− | 286.64 | 284.23 |
2.2.3 119Sn Mössbauer spectroscopy
The 119Sn Mössbauer spectrum of 11 (78 K data, Fig. 6) shows two distinct signals in an almost perfect 1:1 ratio with isomer shifts (δ) of 3.52(1) mm s−1 and 3.45(1) mm s−1 (Table 4), which clearly indicate the oxidation state +II for tin. Furthermore, the electric quadrupole splitting (ΔEQ) of the two signals differs notably and amounts to 1.28(2) mm s−1 and 1.77(2) mm s−1 demonstrating the unsymmetrical coordination around the tin atom in both cases. These values compare to isomer shifts of 2.94(1) and 3.18(1) mm s−1 with electric quadrupole splittings of 1.63(1) and 1.64(1) mm s−1, respectively, for the ion pair 12 [42]. Concerning the shifts of the ion pair 12, the signal at δ=2.94 mm s−1 can be attributed to the tin(II) atom within the DIMPY complex while the second signal at δ=3.18 mm s−1 can be assigned to the trichlorostannate(II) anion [43], [44]. For comparison, the neutral complex [DIMPY·Sn(0)] exhibits a quadrupole split signal at δ=2.64 mm s−1 with ΔEQ=2.11 mm s−1 at 78 K [42]. The sample shows a small contribution (ca. 1%) of a tetravalent tin species (isomer shifts around 0 mm s−1), most likely due to hydrolyses of the sample and formation of SnO2 [45].
119Sn Mössbauer spectroscopic data at 78 K for the two signals of 11: δ=isomer shift, ΔEQ=electric quadrupole splitting, Γ=experimental line width.
Compound | δ (mm s−1) | ΔEQ (mm s−1) | Γ (mm s−1) | Ratio (%) |
---|---|---|---|---|
11 Site A | 3.52(1) | 1.28(2) | 0.86(2) | 49(1) |
11 Site B | 3.45(1) | 1.77(2) | 0.75(1) | 50* |
Impurity | 0.1(1) | 0 | 0.75* | 1(1) |
Parameters marked with an asterisk were kept fixed during the fitting procedure.
3 Summary
The mono- and diimino pyridine ligands SIMPY and DIMPY were reacted with germanium(II) and tin(II) salts EX2 (X=Cl, Ge, I, OTf). While the SIMPY provides the neutral 1:1 complexes [SIMPY·EX2] for all halides and the triflate salts, introduction of a second donor side arm as in the 2,6-substituted pyridine ligand DIMPY gives rise to the ion pairs [DIMPY·EX]+[EX3]− for the halides of germanium and tin. Surprisingly, in the case of the DIMPY complex of tin(II) triflate, yet again a neutral 1:1 complex is isolated, in which two different tin centers exist in the solid state with coordination numbers five and six, where an additional THF donor molecule is coordinated to the neutral complex in the latter species. 119Sn Mössbauer data support the existence of two distinct tin sites in the solid state structure of the ditriflate [DIMPY·SnOTf2], with isomer shifts (δ) of 3.52(1) mm s−1 and 3.45(1) mm s−1, clearly indicating the oxidation state +II for tin. More importantly, the electric quadrupole splitting (ΔEQ) of the two sites differs notably and accounts to 1.28(2) mm s−1 for site 1 and 1.77(2) mm s−1 for site 2.
![Fig. 5: Molecular structures of two crystallographically independent molecules of [DIMPY·SnOTf2], 11, in the solid state, where Sn2 is 5-coordinate (left) and Sn1 is 6-coordinate (right). Ellipsoids are drawn at the 50% probability level, hydrogen atoms are omitted for clarity.](/document/doi/10.1515/znb-2017-0128/asset/graphic/j_znb-2017-0128_fig_005.jpg)
Molecular structures of two crystallographically independent molecules of [DIMPY·SnOTf2], 11, in the solid state, where Sn2 is 5-coordinate (left) and Sn1 is 6-coordinate (right). Ellipsoids are drawn at the 50% probability level, hydrogen atoms are omitted for clarity.
4 Experimental section
4.1 General procedures
All manipulations were carried out using modified Schlenk techniques or in a mBRAUN UNIlab drybox under an atmosphere of dry nitrogen. Solvents were dried using an Innovative Technologies column solvent purification system. Chemicals were purchased from Aldrich company and used as received. 1H, 13C, and 119Sn spectroscopic data was recorded on a Varian Mercury 300 MHz spectrometer (operating at 300.23 MHz for 1H, 75.50 MHz for 13C, 111.92 MHz for 119Sn). NMR spectra were referenced to deuterated benzene, which was dried over potassium metal, and recorded at 25°C. Elemental analysis was performed on a Elementar Vario EL III. Decomposition regions were determined using a Büchi 535 instrument. The ligands SIMPY [33] and DIMPY [13] were synthesized according to literature procedures.
4.2 X-ray crystallography
Suitable single crystals of the compounds were immersed in silicone oil, mounted using a glass fiber and frozen in the cold nitrogen stream (100 K). X-ray diffraction data were collected at low temperature on a Bruker Kappa APEX II diffractometer using Mo Ka radiation (λ=0.71073 Å) generated by an INCOATEC micro-focus source. The data reduction and absorption correction was performed with the Bruker Smart [46] and Sadabs [47] routines, respectively. The structures were solved by Direct Methods with Shelxt [48], [49] and refined with Shelxl [50], [51], [52] by least-squares minimization against F2 using first isotropic and later anisotropic displacement parameters for all non-hydrogen atoms. Hydrogen atoms were added to the structure models on calculated positions using the riding model. The space group assignments and structural solutions were evaluated using Platon [53], [54].
The files CCDC 1557339–1557346 contain the supplementary crystallographic data. These data can be obtained free of charge via www.ccdc.cam.ac.uk/conts/retrieving.html (or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax +44 1223 336033; e-mail deposit@ccdc.cam.ac.uk).
4.3 119Sn Mössbauer spectroscopy
A Ca119mSnO3 source was available for the 119Sn investigations of 11 and a palladium foil of 0.05 mm thickness was used to reduce the tin K X-rays concurrently emitted by this source. The measurement was performed in the usual transmission geometry in a commercial helium flow cryostat at 78 K. The source was kept at room temperature in all experiments. The sample was placed within a thin-walled PMMA container at a thickness corresponding to about 10 mg Sn per cm2. The spectrum was fitted with the Normos-90 program system [55].

Experimental (data points) and simulated (continuous lines) 119Sn Mössbauer spectrum of 11 at 78 K.
4.4 Syntheses
Synthesis of [SIMPY·GeCl2] (1)
GeCl2·dioxane (434 mg, 1.84 mmol, 1 eq) was dissolved in 4 mL THF and the SIMPY ligand 2-[ArN=CH](NC5H4) (Ar=C6H3-2-iPr2) (500 mg, 1.84 mmol, 1 eq) was added at room temperature. Color change was immediate as the solution went from clear transparent to a bright orange. The solvent was removed and the solid then dried under vacuum. Orange needle-like crystals suitable for X-ray analysis were obtained by recrystallization in THF and storing at −30°C. Yield: 99%; m. p.: 153°C. – 1H NMR (300.22 MHz, C6D6): δ=8.84 (d, 1H, PyH), 8.65 (s, 1H, HC=N), 7.97 (d, 1H, PyH), 7.22 (m, 4H, CH arom), 6.83 (t, 1H, PyH), 3.18 (sept, 2H, CHiPr), 1.23 (d, 12H, CH3iPr). – 13C NMR (75.50 MHz, C6D6): δ=163.87 (C=N, imine), 149.98 (ArH), 148.35 (ArH), 140.39 (ArH), 139.24 (ArH), 127.09 (ArH), 125.48 (ArH), 123.93(ArH), 123.10 (ArH), 28.47 (CHiPr), 24.07 (CH3iPr). – Elemental analysis (%) for C18H22Cl2N2Ge (409.92): calcd. C 52.74, H 5.41, N 6.83; found C 51.51, H 5.25, N 6.49.
Synthesis of [SIMPY·GeBr2] (2)
Procedure as described for [SIMPY·GeCl2]. GeBr2·dioxane (241 mg, 0.75 mmol, 1 eq) was dissolved in 4 mL THF and the SIMPY ligand (200 mg, 0.75 mmol, 1 eq) was added at room temperature. Dark orange crystals were obtained after recrystallizing in a THF-benzene mixture (2:1) at room temperature. Yield: 97%. m. p.: 169.9–171.9°C. – 1H NMR (300.22 MHz, C6D6): δ=8.85 (d, 1H, PyH), 8.49 (s, 1H, HC=N), 7.87 (d, 1H, PyH), 7.21 (m, 4H, CH arom), 6.76 (t, 1H, PyH), 3.17 (sept, 2H, CHiPr), 1.22 (d, 12H, CH3iPr). – 13C NMR (75.50 MHz, C6D6): δ=161.57 (C=N, imine), 150.67 (ArH), 148.50 (ArH), 139.54 (ArH), 138.17 (ArH), 126.48 (ArH), 126.15 (ArH), 124.46 (ArH), 123.59 (ArH), 28.28 (CHiPr), 23.73 (CH3iPr). – Elemental analysis (%) for C18H22Br2N2Ge (498.83): calcd. C 43.34, H 4.45, N 5.62; found C 43.17, H 4.39, N 5.68.
Synthesis of [SIMPY·SnCl2] (3)
Procedure as described for [SIMPY·GeCl2]. SnCl2 (338 mg, 1.84 mmol, 1 eq) was dissolved in 4 mL THF and the SIMPY ligand (500 mg, 1.84 mmol, 1 eq) was added at room temperature. Orange crystals were obtained after recrystallizing in THF at −30°C. Yield: 98%; m. p.: 205°C. – 1H NMR (300.22 MHz, D2O capillary/THF)): δ=9.18 (d, 1H, PyH), 8.50 (s, 1H, HC=N), 8.32 (d, 1H, PyH), 8.11 (t, 1H, PyH), 7.68 (t, 1H, PyH), 7.26 (d, 1H, CH arom), 7.22 (t, 1H, CH arom), 3.13 (sept, 2H, CHiPr), 1.27 (d, 12H, CH3iPr). – 13C NMR (75.50 MHz, D2O capillary/THF): δ=163.28 (C=N, imine), 149.94 (ArH), 137.87 (ArH), 137.49 (ArH), 126.09 (ArH), 124.78 (ArH), 123.27 (ArH), 122.97 (ArH), 27.99 (CHiPr), 23.22 (CH3iPr). – 119Sn NMR (112.17 MHz, D2O capillary/THF): δ=−216.3. – Elemental analysis (%) for C18H22Cl2N2Sn (456.00): calcd. C 47.41, H 4.86, N 6.14; found C 45.87, H 4.68, N 5.91.
Synthesis of [SIMPY·SnBr2] (4)
Procedure as described for [SIMPY·GeCl2]. SnBr2 (209 mg, 0.75 mmol, 1 eq) was dissolved in 4 mL THF and the SIMPY ligand (200 mg, 0.75 mmol, 1 eq) was added at room temperature. Red cubic crystals were obtained after recrystallizing in a THF-benzene mixture (2:1) at room temperature. Yield: 84%; m. p.: 208°C. – 1H NMR (300.22 MHz, CDCl3): δ=9.94 (d, 1H, PyH), 8.38 (s, 1H, HC=N), 8.18 (t, 1H, PyH), 7.95 (d, 1H, PyH), 7.74 (t, 1H, PyH), 7.18 (m, 3H, CH arom), 2.94 (sept, 2H, CHiPr), 1.12 (d, 12H, CH3iPr). – 13C NMR (75.50 MHz, CDCl3): δ=160.82 (C=N, imine), 151.77 (ArH), 149.24 (ArH), 141.18 (ArH), 138.85 (ArH), 128.17 (ArH), 126.85 (ArH), 123.84 (ArH), 123.26 (ArH), 28.41 (CHiPr), 24.41 (CH3iPr). – 119Sn NMR (112.17 MHz, D2O capillary/THF): δ=−80.1. – Elemental analysis (%) for C18H22Br2N2Sn (544.90): calcd. C 39.68, H 4.07, N 5.14; found C 39.80, H 4.02, N 5.22.
Synthesis of [SIMPY·SnI2] (5)
Procedure as described for [SIMPYGeCl2]. SnI2 (932 mg, 2.50 mmol, 1 eq) was dissolved in 4 mL THF and the SIMPY ligand (400 mg, 2.50 mmol, 1eq) was added at room temperature. Dark red cubic crystals were obtained after recrystallizing in a THF-benzene mixture (2:1) at room temperature. Yield: 79%; m. p.: 191.4–192.3°C. – 1H NMR (300.22 MHz, C6D6): δ=9.21 (d, 1H, PyH), 8.37 (s, 1H, HC=N), 8.00 (d, 1H, PyH), 7.82 (t, 1H, PyH), 7.36 (t, 1H, PyH), 7.07 (m, 4H, CH arom), 3.00 (sept, 2H, CHiPr), 1.08 (d, 12H, CH3iPr). – 13C NMR (75.50 MHz, C6D6): δ=163.16 (C=N, imine), 151.60 (ArH), 151.18 (ArH), 138.89 (ArH), 137.81 (ArH), 126.58 (ArH), 125.32 (ArH), 124.91 (ArH), 124.49 (ArH), 123.09 (ArH), 27.96 (CHiPr), 23.34 (CH3iPr). – 119Sn NMR (112.17 MHz, D2O capillary/THF): δ=317.2 broad signal. Elemental analysis (%) for C18H22I2N2Sn (638.91): calcd. C 33.84, H 3.41, N 4.38; found C 31.45, H 2.73, N 3.84.
Synthesis of [SIMPY·SnOTf2] (6)
SIMPY ligand (0.5 g, 1.88 mmol, 1 eq) was dissolved in 5 mL THF and SnOTf2 (0.78 g, 1.88 mmol, 1 eq) was added. The light yellow solution turned bright orange. All volatile components were removed in vacuo and a bright orange powder was obtained. Yield: 93%; m. p.: 67.5–68.6°C. – 1H NMR (300.22 MHz, C6D6): δ=9.43 (broad s, 1H, HC=N), 8.19 (broad d, 1H, PyH), 7.58 (broad s, 1H, PyH), 7.18 (m, 3H, CH arom), 2.65 (broad sept, 2H, CHiPr), 1.24 (broad d, 12H, CH3iPr). – 13C NMR (75.50 MHz, C6D6): δ=167.22 (C=N, imine), 150.72 (ArH), 147.77 (ArH), 144.97 (ArH), 142.17 (ArH), 141.02 (ArH), 129.51 (ArH), 124.34 (ArH), 122.29 (ArH), 118.06 (ArH), 28.16 (CHiPr), 23.18 (CH3iPr), 23.14 (CH3iPr). – 119Sn NMR (112.17 MHz, C6D6-THF): δ=−27.8. Elemental analysis (%) for C19H22F3N2O9S3Sn (694.28): calcd. C 32.87, H 3.19, N 4.03; found C 32,58; H 3.23; N 3.99.
Synthesis of [DIMPY·GeCl]+[GeCl3]– (7)
GeCl2·dioxane (0.51 g, 2.20 mmol, 2 eq) was dissolved in 5 mL THF and DIMPY (0.50 g, 1.10 mmol, 1 eq) was added at room temperature. Color change was immediate as the solution went from clear transparent to orange. The solvent was reduced and the solid then dried under vacuum. Orange needle-like crystals suitable for X-ray analysis were obtained by recrystallization in a mixture of THF-benzene at room temperature. Yield: 98%; m. p.: 158.6–160.7°C. 1H NMR (300.22 MHz, C6D6): δ=8.61(s, 1H, HC=N), 8.55 (s, 1H, HC=N), 8.44 (t, 1H, PyH), 8.30 (td, 1H, PyH), 7.15 (m, 6H, CH arom), 3.15 (broad sept, 4H, CHiPr), 1.25 (broad d, 12H, CH3iPr), 1.16 (broad d, 12H, CH3iPr). – 13C NMR (75.50 MHz, C6D6): δ 162.87 (C=N, imine), 159.75 (C=N, imine), 154.67 (ArH), 148.66 (ArH), 146.34 (ArH), 139.57 (ArH), 137.13 (ArH), 136.19 (ArH), 132.67 (ArH), 128.18 (ArH), 124.77 (ArH), 123.15 (ArH), 122.19 (ArH), 28.08 (CHiPr), 23.14 (CH3iPr). – Elemental analysis (%) for C31H39Cl4N3Ge2 (740.75): calcd. C 50.26, H 5.31, N 5.67; found C 51.47, H 6.66, N 7.03.
Synthesis of [DIMPY·GeBr]+[GeBr3]– (8)
Procedure as described for [DIMPYGeCl]+[GeCl3]− (7). GeBr2·dioxane (0.28 g, 0.88 mmol, 2 eq) was dissolved in 5 mL THF and DIMPY (0.20 g, 0.44 mmol, 1 eq) was added at room temperature. Orange needle like crystals suitable for X-ray analysis were obtained by recrystallization in a mixture of THF-benzene at room temperature. Yield: 98%; m. p.: 171.7–172.5°C. – 1H NMR (300.22 MHz, C6D6): δ=8.61(s, 1H, HC=N), 8.55 (s, 1H, HC=N), 8.44 (t, 1H, PyH), 8.30 (td, 1H, PyH), 7.15 (m, 6H, CH arom), 3.15 (broad sept, 4H, CHiPr), 1.25 (broad d, 12H, CH3iPr), 1.16 (broad d, 12H, CH3iPr). – 13C NMR (75.50 MHz, C6D6-THF): δ 160.06 (C=N, imine), 148.41 (ArH), 146.34 (ArH), 141.71 (ArH), 140.28 (ArH), 139.58 (ArH), 132.43 (ArH), 129.45 (ArH), 128.16 (ArH), 124.48 (ArH), 124.27 (ArH), 30.02 (CHiPr), 28.64 (CH3iPr). – Elemental analysis (%) for C31H39Cl4N3Ge2 (740.75): calcd. C 50.26, H 5.31, N 5.67; found C 50.33, H 4.59, N 4.44.
Synthesis of [DIMPY·SnBr]+[SnBr3]– (9)
SnBr2 (0.61 g, 2.20 mmol, 2 eq) was dissolved in 5 mL THF and DIMPY (0.50 g, 1.10 mmol, 1 eq) was added at room temperature resulting in an immediate color change from pale yellow to orange. The solvent was removed and the solid then dried under vacuum. Dark orange needles were obtained after recrystallizing in a THF-benzene mixture (2:1) at room temperature. Yield: 78%; m. p.: 196.6–197.8°C. – 1H NMR (300.23 MHz, C6D6): δ=8.61 (s, 1H, HC=N), 8.48 (s, 1H, HC=N), 8.38 (d, 1H, PyH), 8.18 (t, 1H, PyH), 7.92 (d, 1H, PyH), 7.24 (m, 6H, CH arom.), 3.21 (broad sept, 4H, CHiPr), 1.33 (d, 12H, CH3iPr), 1.23 (d, 12H, CH3iPr) ppm. – 13C NMR (75.50 MHz, C6D6): δ=162.83 (C=N, imine), 159.39 (C=N, imine), 154.63 (Ar-CH), 148.12 (Ar-CH), 146.08 (Ar-CH), 140.30 (Ar-CH), 139.58 (Ar-CH), 137.00 (Ar-CH), 132.09 (Ar-CH), 128.31 (Ar-CH), 124.78 (Ar-CH), 124.37 (Ar-CH), 123.15 (Ar-CH), 122.23 (Ar-CH), 28.07 (CHiPr), 25.40 (CH3iPr), 23.13 (CH3iPr) ppm. – 119Sn NMR (111.92 MHz, C6D6): δ=98.2 (DimpySnBr), −405.8 (SnBr3− anion) ppm. – Elemental analysis (%) for C31H39Br4N3Sn2·4THF (1299.14): calcd. C 43.45, H 5.51, N 3.23; found: C 43.28, H 5.29, N 3.31.
Synthesis of [DIMPY·SnI]+[SnI3]– (10)
Procedure as described for [DIMPYGeCl]+[GeCl3]− (7). SnI2 (0.82 g, 2.20 mmol, 2 eq) was dissolved in 5 mL THF and the ligand DIMPY (0.50 g, 1.10 mmol, 1 eq) was added at room temperature. Yield: 91%; m. p.: 196.6–197.8°C. – 1H NMR (300.23 MHz, C6D6): δ=9.24 (d, 1H, PyH), 8.40 (s, 1H, HC=N), 8.04 (d, 1H, PyH), 7.86 (t, 1H, PyH), 7.38 (t, 2H, CH arom.), 7.08 (m, 4H, CH arom.), 3.02 (broad sept, 4H, CHiPr), 1.09 (d, 24H, CH3iPr) ppm. – 13C NMR (75.50 MHz, C6D6): δ=163.30 (C=N, imine), 162.82 (Ar-CH), 159.36 (Ar-CH), 151.03 (Ar-CH), 146.00 (Ar-CH), 139.58 (Ar-CH), 137.88 (Ar-CH), 132.76 (Ar-CH), 127.97 (Ar-CH), 126.87 (Ar-CH), 125.49 (Ar-CH), 123.16 (Ar-CH), 122.59 (Ar-CH), 27.98 (CHiPr), 23.45 (CH3iPr) ppm. – Elemental analysis (%) for C33H47I4N3Sn2 (1230.80): calcd. C 32.20, H 3.85, N 3.41; found C 33.61, H 3.77, N 4.33.
Synthesis of [DIMPY·SnOTf2] (11)
DIMPY ligand (0.5 g, 1.10 mmol, 1 eq) was dissolved in 5 mL THF and SnOTf2 (0.46 g, 1.10 mmol, 1 eq) was added. The solvent was reduced in vacuo and an orange powder was obtained. Cubic orange crystals were obtained via recrystallizing in THF-benzene mixture at room temperature. Yield: 78%; m. p.: 202.7–204.9°C. – 1H NMR (300.23 MHz, C6D6): δ=8.18 (s, 2H, HC=N), 7.31 (t, 1H, PyH), 7.14 (m, 8H, CH arom.), 3.52 (broad sept, 4H, CHiPr), 1.27 (d, 24H, CH3iPr) ppm. – 13C NMR (75.50 MHz, C6D6): δ=163.48 (C=N, imine), 151.07 (Ar-CH), 144.48 (Ar-CH), 141.53 (Ar-CH), 131.03 (Ar-CH), 128.46 (Ar-CH), 124.79 (Ar-CH), 122.28 (Ar-CH), 118.05 (Ar-CH), 27.62 (CHiPr), 24.65 (CH3iPr) ppm. – 119Sn NMR (111.92 MHz, C6D6): δ=−46.4 ppm. – Elemental analysis (%) for C39H57F6N3O6S2Sn (960.72): calcd. C 48.76, H 5.98, N 4.37; found C 47.01, H 4.93 N 4.42.
Acknowledgements
The authors thank Barbara Seibt and Monika Filzwieser for performing the elemental analyses. Astrid Falk is gratefully acknowledged for support with the synthetic work. The NAWI Graz Project is gratefully acknowledged for providing infrastructure.
References
[1] S. C. Bart, E. Lobkovsky, P. J. Chirik, J. Am. Chem. Soc.2004, 126, 13794.10.1021/ja046753tSearch in Google Scholar
[2] S. C. Bart, K. Chłopek, E. Bill, M. W. Bouwkamp, E. Lobkovsky, F. Neese, K. Wieghardt, P. J. Chirik, J. Am. Chem. Soc.2006, 128, 13901.10.1021/ja064557bSearch in Google Scholar
[3] S. C. E. Stieber, C. Milsmann, J. M. Hoyt, Z. R. Turner, K. D. Finkelstein, K. Wieghardt, S. DeBeer, P. J. Chirik, Inorg. Chem.2012, 51, 3770.10.1021/ic202750nSearch in Google Scholar
[4] S. C. Bart, E. Lobkovsky, E. Bill, K. Wieghardt, P. J. Chirik, Inorg. Chem.2007, 46, 7055.10.1021/ic700869hSearch in Google Scholar
[5] J. M. Darmon, Z. R. Turner, E. Lobkovsky, P. J. Chirik, Organometallics2012, 31, 2275.10.1021/om201212mSearch in Google Scholar
[6] G. J. P. Britovsek, M. Bruce, V. C. Gibson, B. S. Kimberley, P. J. Maddox, S. Mastroianni, S. J. McTavish, C. Redshaw, G. A. Solan, S. Stroemberg, J. Am.Chem. Soc.1999, 121, 8728.10.1021/ja990449wSearch in Google Scholar
[7] J. H. Docherty, J. Peng, A. P. Dominey, S. P. Thomas, Nature Chem.2017, 9, 595.10.1038/nchem.2697Search in Google Scholar
[8] D. Enright, S. Gambarotta, G. P. A. Yap, P. H. M. Budzelaar, Angew. Chem. Int. Ed.2002, 41, 3873.10.1002/1521-3773(20021018)41:20<3873::AID-ANIE3873>3.0.CO;2-8Search in Google Scholar
[9] H. P. Nayek, N. Arleth, I. Trapp, M. Loeble, P. Ona-Burgos, M. Kuzdrowska, Y. Lan, A. K. Powell, F. Breher, P. W. Roesky, Chem. Eur. J.2011, 17, 10814.10.1002/chem.201101646Search in Google Scholar
[10] M. Arrowsmith, M. S. Hill, G. Kociok-Köhn, Organometallics2010, 29, 4203.10.1021/om100649zSearch in Google Scholar
[11] I. J. Blackmore, V. C. Gibson, P. B. Hitchcock, C. W. Rees, D. J. Williams, A. J. P. White, J. Am. Chem. Soc.,2005, 127, 6012.10.1021/ja042657gSearch in Google Scholar PubMed
[12] T. Summerscales, T. W. Myers, L. A. Berben, Organometallics2012, 31, 3463.10.1021/om300242qSearch in Google Scholar
[13] J. Scott, S. Gambarotta, I. Korobkov, Q. Knijnenburg, B. de Bruin, P. H. M. Budzelaar, J. Am. Chem. Soc.2005, 127, 17204.10.1021/ja056135sSearch in Google Scholar PubMed
[14] Q. Knijnenburg, J. M. M. Smits, P. H. M. Budzelaar, Organometallics2006, 25, 1036.10.1021/om050936mSearch in Google Scholar
[15] M. Bruce, V. C. Gibson, C. Redshaw, G. A. Solan, A. J. P. White, D. J. Williams, Chem. Commun. 1998, 2523.10.1039/a807950aSearch in Google Scholar
[16] T. Jurca, J. Lummiss, T. J. Burchell, S. I. Gorelsky, D. S. Richeson, J. Am. Chem. Soc.2009, 131, 4608.10.1021/ja901128qSearch in Google Scholar PubMed
[17] R. J. Baker, C. Jones, M. Kloth, D. P. Mills, New J. Chem.2004, 28, 207.10.1039/b310592jSearch in Google Scholar
[18] T. W. Myers, L. A. Berben, J. Am. Chem. Soc. 2011, 133, 11865.10.1021/ja203842sSearch in Google Scholar PubMed
[19] T. W. Myers L. A. Berben, Inorg. Chem. 2012, 51, 1480.10.1021/ic201729bSearch in Google Scholar PubMed
[20] K. Kowolik, M. Shanmugam,T. W. Myers, C. D. Cates, L. A. Berben, Dalton Trans.2012, 41, 7969.10.1039/c2dt30112aSearch in Google Scholar PubMed
[21] T. W. Myers, N. Kazem, S. Stoll, D. R. Britt, M. Shanmugam, L. A. Berben, J. Am. Chem. Soc. 2011, 133, 8662.10.1021/ja2015718Search in Google Scholar PubMed
[22] T. W. Myers, L. A. Berben, Inorg. Chem. 2012, 51, 1480.10.1021/ic201729bSearch in Google Scholar PubMed
[23] T. W. Myers, A. L. Holmes, L. A. Berben, Inorg. Chem. 2012, 51, 8997.10.1021/ic301128mSearch in Google Scholar PubMed
[24] C. D. Cates, T. W. Myers, L. A. Berben, Inorg. Chem. 2012, 51, 11891.10.1021/ic301792wSearch in Google Scholar PubMed
[25] T. W. Myers, G. M. Yee, L. A. Berben, Eur. J. Inorg. Chem. 2013, 3831.10.1002/ejic.201300192Search in Google Scholar
[26] T. W. Myers, L. A. Berben, Chem. Commun.2013, 49, 4175.10.1039/C2CC37208HSearch in Google Scholar PubMed
[27] M. Shen, W. Zhang, K. Nomura, W.-H. Sun, Dalton Trans. 2009, 9000.10.1039/b910155aSearch in Google Scholar PubMed
[28] J. S. Casas, E. E. Castellano, J. Ellena, M. S. Garcìa-Tasende, A. Sànchez, J. Sordo, E. M. Vazquez-Lòpez, M. J. Vidarte, Z. Anorg. Allg. Chem.2003, 629, 261.10.1002/zaac.200390042Search in Google Scholar
[29] C. D. Martin, P. J. Ragogna, Dalton Trans.2011, 40, 11976.10.1039/c1dt11111fSearch in Google Scholar PubMed
[30] G. Reeske, A. H. Cowly, Chem. Commun.2006, 4856.10.1039/B610491FSearch in Google Scholar PubMed
[31] K. Nomiya, M. Sekino, A. Ishikawa, M. Honda, N. C. Yokoyama, K. Kasuga, H. Yokoyama, S. Nakano, K. J. Onodera, Inorg. Biochem.2004, 98, 601.10.1016/j.jinorgbio.2004.01.011Search in Google Scholar PubMed
[32] P. A. Rupar, V. N. Staroverov, K. M. Baines, Science, 2008, 322, 1360.10.1126/science.1163033Search in Google Scholar PubMed
[33] Z. Benko, S. Burck, D. Gudat, M. Nieger, L. Nyulaszi, N. Shore, Dalton Trans.2008, 4937.10.1039/b804621bSearch in Google Scholar PubMed
[34] G. Reeske, A. H. Cowly, Chem. Commun.2006, 1784.10.1039/b602017hSearch in Google Scholar PubMed
[35] C. D. Martin, C. M. Le, P. J. Ragogna, J. Am. Chem. Soc.2009, 131, 15126.10.1021/ja9073968Search in Google Scholar PubMed
[36] A. P. Singh, H. W. Roesky, E. Carl, D. Stalke, J.-P. Demers, A. Lange, J. Am. Chem. Soc.2012, 134, 4998.10.1021/ja300563gSearch in Google Scholar PubMed
[37] T. Jurca, L. K. Hiscock, I. Korobkov, C. N. Rowley, D. S. Richeson, Dalton Trans. 2014, 43, 690.10.1039/C3DT52227JSearch in Google Scholar
[38] M. Bouska, L. Dostal, A. Ruzicka, R. Jambor, Organometallics2013, 32, 1995.10.1021/om400082tSearch in Google Scholar
[39] E. Magdzinski, P. Gobbo, M. S. Workentin, P. J. Ragogna, Inorg. Chem. 2013, 52, 11311.10.1021/ic401598eSearch in Google Scholar PubMed
[40] M. Novak, M. Bouska, L. Dostal, M. Lutter, K. Jurkschat, J. Turek, F. de Proft, Z. Ruzickova, R. Jambor, Eur. J. Inorg. Chem. 2017, 1292.10.1002/ejic.201700098Search in Google Scholar
[41] M. Bouska, L. Dostal, M. Lutter, B. Glowacki, Z. Ruzickova, D. Beck, R. Jambor, K. Jurkschat, Inorg. Chem. 2015, 54, 6792.10.1021/acs.inorgchem.5b00678Search in Google Scholar
[42] J. Flock, A. Suljanovic, A. Torvisco, W. Schöfberger, B. Gerke, R. Pöttgen, R. C. Fischer, M. Flock, Chem. Eur. J. 2013, 19, 15504.10.1002/chem.201301340Search in Google Scholar
[43] J. K. Stalick, D. M. Meek, B. Y. K. Ho, J. J. Zuckerman, J. Chem. Soc. Chem. Commun. 1972, 630.10.1039/C39720000630Search in Google Scholar
[44] T. Kégl, L. Kollár, G. Szalontai, E. Kuzmann, A. Vértes, J. Organomet. Chem. 1996, 507, 75.10.1016/0022-328X(95)05630-8Search in Google Scholar
[45] P. E. Lippens, Phys. Rev. B1999, 60, 4576.10.1103/PhysRevB.60.4576Search in Google Scholar
[46] Smart, Bruker AXS Inc., Madison, Wisconsin (USA) 2007.Search in Google Scholar
[47] Sadabs, Bruker AXS Inc., Madison, Wisconsin (USA) 2001.Search in Google Scholar
[48] G. M. Sheldrick, Shelxt, Integrated Space-Group and Crystal-Structure Determination, University of Göttingen, Göttingen (Germany) 2015.10.1107/S2053273314026370Search in Google Scholar PubMed PubMed Central
[49] G. M. Sheldrick, Acta Crystallogr.2015, A71, 3.10.1107/S2053273314026370Search in Google Scholar
[50] G. M. Sheldrick, Shelxl-97, Program for the Refinement of Crystal Structures, University of Göttingen, Göttingen (Germany) 1997.Search in Google Scholar
[51] G. M. Sheldrick, Acta Crystallogr.2008, A64, 112.10.1107/S0108767307043930Search in Google Scholar PubMed
[52] G. M. Sheldrick, Acta Crystallogr.2015, C71, 3.Search in Google Scholar
[53] A. L. Spek, J. Appl. Crystallogr.2003, 36, 7.10.1107/S0021889802022112Search in Google Scholar
[54] A. L. Spek, Acta Crystallogr.2009, D65, 148.10.1107/S090744490804362XSearch in Google Scholar
[55] R. A. Brand, Normos, Mössbauer fitting Program, Universität Duisburg, Duisburg (Germany) 2002.Search in Google Scholar
©2017 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- In this Issue
- Preface
- Congratulations to Dietrich Gudat
- On the dimorphism of Pr6Mo10O39
- Rhodium-rich silicides RERh6Si4 (RE=La, Nd, Tb, Dy, Er, Yb)
- Coordination of the ambiphilic phosphinoborane tBu2PCH2BPh2 to Cu(I)Cl
- N-Heterocyclic germylenes and stannylenes of the type [Fe{(η5-C5H4)NR}2E] with bulky alkyl substituents
- Die Europium(II)-Oxidhalogenide Eu2OBr2 und Eu2OI2
- Structure and spectroscopic properties of porphyrinato group 14 derivatives: Part I – Phenylacetylido ligands
- Synthesis, solid-state structures and reduction reactions of heteroleptic Ge(II) and Sn(II) β-ketoiminate complexes
- Reactions of Al/P, Ga/P and P–H functionalized frustrated Lewis pairs with azides and a diazomethane – formation of adducts and capture of nitrenes
- Metal carbonyl complexes of potentially ambidentate 2,1,3-benzothiadiazole and 2,1,3-benzoselenadiazole acceptors
- Lithium alkaline earth tetrelides of the type Li2AeTt (Ae=Ca, Ba, Tt=Si, Ge, Sn, Pb): synthesis, crystal structures and physical properties
- Magnetic properties of the germanides RE3Pt4Ge6 (RE=Y, Pr, Nd, Sm, Gd–Dy)
- Overcrowded aminophospanitrenes: a case study
- PCl bond length depression upon coordination of a diazaphosphasiletidine to a group 13 element Lewis acid or a transition metal carbonyl fragment – Synthesis and structural characterization of diazaphosphasiletidine adducts with P-coordination
- Iminopyridine ligand complexes of group 14 dihalides and ditriflates – neutral chelates and ion pair formation
- On the structure of the P-iodo-, bromo- and chloro-bis(imino)phosphoranes: A DFT study
- (Dicyclohexyl(2-(dimesitylboryl)phenyl)phosphine: en route to stable frustrated Lewis pairs-hydrogen adducts in water
- Insertion of phenyl isocyanate into mono- and diaminosilanes
Articles in the same Issue
- Frontmatter
- In this Issue
- Preface
- Congratulations to Dietrich Gudat
- On the dimorphism of Pr6Mo10O39
- Rhodium-rich silicides RERh6Si4 (RE=La, Nd, Tb, Dy, Er, Yb)
- Coordination of the ambiphilic phosphinoborane tBu2PCH2BPh2 to Cu(I)Cl
- N-Heterocyclic germylenes and stannylenes of the type [Fe{(η5-C5H4)NR}2E] with bulky alkyl substituents
- Die Europium(II)-Oxidhalogenide Eu2OBr2 und Eu2OI2
- Structure and spectroscopic properties of porphyrinato group 14 derivatives: Part I – Phenylacetylido ligands
- Synthesis, solid-state structures and reduction reactions of heteroleptic Ge(II) and Sn(II) β-ketoiminate complexes
- Reactions of Al/P, Ga/P and P–H functionalized frustrated Lewis pairs with azides and a diazomethane – formation of adducts and capture of nitrenes
- Metal carbonyl complexes of potentially ambidentate 2,1,3-benzothiadiazole and 2,1,3-benzoselenadiazole acceptors
- Lithium alkaline earth tetrelides of the type Li2AeTt (Ae=Ca, Ba, Tt=Si, Ge, Sn, Pb): synthesis, crystal structures and physical properties
- Magnetic properties of the germanides RE3Pt4Ge6 (RE=Y, Pr, Nd, Sm, Gd–Dy)
- Overcrowded aminophospanitrenes: a case study
- PCl bond length depression upon coordination of a diazaphosphasiletidine to a group 13 element Lewis acid or a transition metal carbonyl fragment – Synthesis and structural characterization of diazaphosphasiletidine adducts with P-coordination
- Iminopyridine ligand complexes of group 14 dihalides and ditriflates – neutral chelates and ion pair formation
- On the structure of the P-iodo-, bromo- and chloro-bis(imino)phosphoranes: A DFT study
- (Dicyclohexyl(2-(dimesitylboryl)phenyl)phosphine: en route to stable frustrated Lewis pairs-hydrogen adducts in water
- Insertion of phenyl isocyanate into mono- and diaminosilanes