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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

[SnF(bipy)(H2O)]2[SnF6], a mixed-valent inorganic tin(II)–tin(IV) compound

crossmark logo

aChemistry, Osnabrück University, Barbarastr. 7, 49069 Osnabrück, Germany
*Correspondence e-mail: hreuter@uos.de

Edited by S.-L. Zheng, Harvard University, USA (Received 8 July 2024; accepted 25 July 2024; online 6 August 2024)

In the title compound, bis­[aqua­(2,2′-bi­pyridine)­fluorido­tin(II)] hexa­fluorido­tin(IV), [SnF(C10H8N2)(H2O)]2[SnF6], an ionic mixed-valent tin(II)–tin(IV) compound, the bivalent tin atom is the center atom of the cation and the tetra­valent tin atom is the center atom of the anion. With respect to the first coordination sphere, the cation is monomeric, with the tin(II) atom having a fourfold seesaw coordination with a fluorine atom in an equatorial position, a water mol­ecule in an axial position and the two nitro­gen atoms of the chelating 2,2′-bi­pyridine ligand in the remaining axial and equatorial positions. The bond lengths and angles of this hypervalent first coordination sphere are described by 2c–2e and 3c–4e bonds, respectively, all of which are based on the orthogonal 5p orbitals of the tin atom. In the second coordination sphere, which is based on an additional, very long tin–fluorine bond that leads to dimerization of the cation, the tin atom is trapezoidal–pyramidally coordinated. The tetra­valent tin atom of the centrosymmetric anion has an octa­hedral coordination. The differences in its tin–fluorine bond lengths are attributed to hydrogen bonding, as the two of the four fluorine atoms are each involved in two hydrogen bonds, linking anions and cations together to form strands.

1. Chemical context

Mixed-valent tin(II)–tin(IV) compounds are most often discovered by chance as a by-product of reactions in which tin(II) compounds are used as starting compounds. Two processes are discussed as potential sources of tetra­valent tin (Gurnani et al., 2013[Gurnani, C., Hector, A. L., Jager, E., Levason, W., Pugh, D. & Reid, G. (2013). Dalton Trans. 42, 8364-8374.]): the oxidation of divalent tin by atmospheric oxygen and the disproportionation of tin(II) into elemental tin and tin(IV). However, there is no evidence that the reaction proceeds in favor of either process, as the amount of the mixed-valent compound is usually limited to a few crystals. For these reasons, elemental tin is often added to the reaction mixture or an inert gas atmosphere is used.

[Scheme 1]

In the case of the title compound, [SnF(H2O)(bipy)]2[SnF6], in a micro-scale experiment in air (and in the absence of elemental tin), we reacted tin difluoride (SnF2) with 2,2′-bi­pyridine (bipy) using N,N-di­methyl­formamide, DMF, as partial solvent and reaction mediator.

2. Structural commentary

The title compound crystallizes in the triclinic space group P[\overline{1}] with half a formula unit in the asymmetric unit resulting in a centrosymmetric [SnIVF6]2− anion and a [SnIIF(H2O)(bipy)]+ cation in a general position (Fig. 1[link]).

[Figure 1]
Figure 1
Ball-and-stick model of the ions found in the crystal of the title compound, showing the atom numbering of the asymmetric unit. With the exception of the H atoms, which are shown as spheres of arbitrary radius, all other atoms are drawn with displacement ellipsoids at the 50% probability level.

In the cation, the first coordination sphere of the bivalent tin atom (Fig. 2[link]) consists of one fluorine atom, the oxygen atom of the water mol­ecule and both nitro­gen atoms of the 2,2′-bi­pyridine ligand. In this seesaw coordination, the fluorine atom [d(Sn—F) = 2.022 (1) Å] adopts an equatorial (eq) position and the water mol­ecule [d(Sn—O = 2.408 (2) Å] an axial (ax) one, while the nitro­gen atoms of the chelating 2,2′-bi­pyridine ligand occupy the two other axial and equatorial positions. According with this arrangement, both tin–nitro­gen distances differ [0.025 (1) Å] with the shorter one to Neq [2.344 (2) Å] and the longer one to Nax [2.369 (2) Å].

[Figure 2]
Figure 2
Ball-and-stick model (left: side view; right: front view) of the the first, seesaw coordination sphere of the bivalent tin atom Sn2 of the cation, highlighting selected bond lengths (Å) and angles (°). With the exception of the H atoms, which are shown as spheres of arbitrary radius, all other atoms are drawn with displacement ellipsoids at the 50% probability level. For clarity, only the carbon atoms between the two nitro­gen atoms of the bi­pyridine ligand are shown, the position of all other carbon atoms are indicated as shortened sticks. Axial bonds are drawn as sticks of reduced thickness in order to underline the presence of a 3c–4e bond.

Within the equatorial plane, the Neq—Sn—F bond angle is 91.5 (1)°, whereas the axis is clearly bent [136.2 (1)°]. Among the axial-to-equatorial bond angles, the bond angle between the nitro­gen atoms is very acute [68.5 (1)°] due to the steric restrictions of the ligand. Conspicuously, the bond angle trans to this chelating bond angle is also quite acute [73.4 (1)°] while the other two are much more obtuse [78.1 (1)°, 79.6 (1)°].

In the VB concept with its localized 2c–2e bonds, the bond ratios within the cation are difficult to describe because the tin atom with its non-bonding 5s electron pair and the eight electrons of the four donor atoms (F,O,N,N) exceeds the electron octet of a main group element. In contrast, the MO theory according to Pimentel and Rundle (Pimentel, 1951[Pimentel, G. C. (1951). J. Chem. Phys. 19, 446-448.]; Rundle, 1949[Rundle, R. E. (1949). J. Chem. Phys. 17, 671-675.]) provides a simple and logical explanation for this hyper-valence in the form of a 3c–4e bond with the exclusive participation of p orbitals on all three atoms. In a seesaw-shaped coordination geometry, this 3c–4e bond is usually found in the axially arranged atoms, while the atoms in the equatorial plane are bonded via 2c–2e bonds through the other p orbitals of the tin atom. This bonding concept not only explains the different tin–nitro­gen bond lengths, but also the bond angles around or smaller than 90° as a result of the orthogonality of the tin p-orbitals. It also explains the remarkable long [2.408 (2) Å] tin–oxygen distance to the water mol­ecule, for which bond lengths between 2.207 (2) Å and 2.226 (2) Å [mean value: 2.226 (17) Å] are observed (Kleeberg et al., 2022[Kleeberg, F. M., Zimmermann, L. W. & Schleid, T. (2022). J. Clust Sci. 33, 2489-2497.]) in electron-precise tin(II) compounds of the type [Sn(OH2)3]2+ with the tin atom at the apex of a trigonal pyramid and the three oxygen atoms of the water mol­ecules at its base.

If only the first coordination sphere is taken as a basis, the cation is monomeric. However, there is a second monomer in its immediate vicinity, whereby both are in contact with each other via a long tin–fluorine bridge [d(Sn—F) = 2.763 (1) Å] resulting from a center of symmetry (Fig. 3[link]). This additional, weak bond is in the trans position [〈(N—Sn—F) = 148.4 (1)°] to the original, equatorial tin–nitro­gen bond and extends the seesaw-shaped, fourfold coordination of the two tin atoms to fivefold, pyramidal coordinations, in which the tin atoms are each below [Δleast-squares = −0.6035 (1) Å] the trapezoidal, strongly uneven [±Δmax, least-squares = −0.147 (1)/0.138 (1) Å] base while the shortest bonds [d(Sn—F) = 2.022 (1) Å] point in the direction of the apical fluorine atoms. This dimerization leads to a bridging angle of 111.8 (1)° at the fluorine atoms and a four-membered, centrosymmetric and therefore exactly planar tin–fluorine ring. Its rhomboidal shape is characterized by acute [68.2 (1)°] angles at the tin atoms and obtuse [111.8 (1)°] ones at the fluorine atoms. Extending the Pimentel–Rundle concept, Musher (1969[Musher, J. I. (1969). Angew. Chem. Int. Ed. Engl. 8, 54-68.]) suggests that such additional, very weak bonds result from a p orbital of the central atom, that is involved in a classical 2c–2e bond on the one hand and a 3c–4e bond on the other. As result, in such ‘asymmetric’ 3c–4e bonds, one ligand is much more strongly bound to the central atom than the trans-ligand. Despite this additional bond, the coordination sphere of the tin atom (Fig. 4[link]) remains hemi-directed (Shimoni-Livny et al., 1998[Shimoni-Livny, L., Glusker, J. P. & Bock, C. W. (1998). Inorg. Chem. 37, 1853-1867.]).

[Figure 3]
Figure 3
Ball-and-stick model of the second, trapezoid-pyramidal coordination sphere of Sn2, highlighting selected bond lengths (Å) and angles (°), viewed down the fluorine–tin bond. With the exception of the H atoms, which are shown as spheres of arbitrary radius, all other atoms are drawn with displacement ellipsoids at the 50% probability level. For clarity, only the carbon atoms between the two nitro­gen atoms of the bi­pyridine ligand are shown, the position of all other carbon atoms are indicated as shortened sticks. Axial bonds are drawn as sticks of reduced thickness in order to underline the presence of a 3c–4e bond, as is the additional asymmetric 3c–4e bond shown with a dashed line. Symmetry code: (i) −x, −y, 1 − z.
[Figure 4]
Figure 4
Space-filling model of the [SnF(bipy)(H2O)]+ cation looking down the apical tin–fluorine bond and visualizing the hemi-directed coordination of the bivalent tin atom Sn2. Atoms whose spheres are penetrated by other atoms are visualized as truncated two-colored spheres, the hydrogen bond is indicated by a dashed red line. Atom colors and van der Waals radii (Å) are as follows: F = green/1.47, H = white/1.10, C = gray/1.70, O = red/1.52, N = blue/1.55 and Sn = brass/2.17. Symmetry code: (i) −x, −y, 1 − z.

The octa­hedral shape of the centrosymmetric [SnF6]2− anion is reflected in bond angles around 90° [88.4 (1)–91.6 (1)°] and very similar tin–fluorine bond lengths. While four of the bonds (F12, F13) are almost identical in length [1.949 (1)/1.942 (2) Å], two bonds (F11) are somewhat longer [1.993 (1) Å] because the fluorine atoms undergo hydrogen bonding with the hydrogen atoms of two different water mol­ecules (Fig. 5[link]).

[Figure 5]
Figure 5
Polyhedron and ball-and-stick model of the octa­hedral, centrosymmetric [SnF6]2− anion with bond lengths (Å) and hydrogen bonds indicated by dashed red lines. All atoms are drawn with displacement ellipsoids at the 50% probability level.

3. Supra­molecular features

Similar to the expansion of the coordination spheres of the bivalent tin atoms through further, asymmetric 3c–4e bonds, the hydrogen bonds between the water mol­ecules of the cation and the fluorine atoms of the anion play a central role in the formation of the crystal structure. Both hydrogen atoms of the water mol­ecule but only one fluorine atom of the anion are involved in these hydrogen bonds whereby a centrosymmetric, almost planar, eight-membered –O—H⋯F ring is formed (Fig. 6[link]). The geometries of the two crystallographic independent hydrogen bonds are given in Table 1[link]. The out-of-plane deflections of the H atoms are +0.011 Å for H101 and −0.001 Å for H102. As a result of these hydrogen bonds, the anions and cations are arranged into strands that expand in the a-axis direction (Fig. 7[link]).

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O1—H101⋯F11i 0.96 1.75 2.705 (2) 170
O1—H102⋯F11ii 0.96 1.75 2.710 (2) 174
Symmetry codes: (i) [-x, -y+1, -z+1]; (ii) [x, y-1, z-1].
[Figure 6]
Figure 6
Ball-and-stick model showing in detail the hydrogen-bonding scheme between cation and anion (dashed red lines). With the exception of the H atoms, which are shown as spheres of arbitrary radius, all other atoms are drawn with displacement ellipsoids at the 50% probability level. The out-of-plane deflections (Å) of the hydrogen atoms are given in square brackets, i = center of symmetry. Symmetry codes: (i) −x, −y + 1, −z + 1; (ii) x, y, z − 1.
[Figure 7]
Figure 7
Stick model of the crystal packing looking down the crystallographic c axis and showing the strands resulting from the hydrogen bonds (dashed red lines) between anions and cations. Color code as in Fig. 4[link].

4. Database survey

The [SnF6]2− anion is often a component of salts with organic and inorganic cations. While in combination with inorganic cations the tin–fluorine distances are often strongly influenced by the cation–anion inter­actions [i.e. Na2SnF6: d(Sn—F) = 1.958 Å; Benner & Hoppe, 1990[Benner, G. & Hoppe, R. (1990). J. Fluor. Chem. 48, 219-227.]], very similar bond lengths [dmean(Sn—F) without/with hydrogen bonds: 1.938 (2)/1.956 (2) Å, 24/16 values] to the title compound are found in combination with organic cations (Lermontov et al., 2010[Lermontov, S. A., Malkova, A. N., Lermontova, E. Kh. & Churakov, A. V. (2010). Russ. J. Gen. Chem. 80, 1825-1830.]; Bouacida et al., 2005[Bouacida, S., Merazig, H., Beghidja, A. & Beghidja, C. (2005). Acta Cryst. E61, m577-m579.]; Taha et al., 1992[Taha, A., Liautard, B. & Granier, W. (1992). Acta Cryst. C48, 1929-1931.]; Cortijo et al., 2017[Cortijo, M., Bulicanu, V., Pedersen, K. S., Rouzières, M., Bendix, J., Clérac, R. & Hillard, E. (2017). Eur. J. Inorg. Chem. pp. 320-325.]; Kokunov et al., 2007[Kokunov, Yu. V., Gorbunova, Yu. E. & Razgonyaeva, G. A. (2007). Russ. J. Inorg. Chem. 52, 158-164.]; Jung et al., 2023[Jung, A., Li, Y. & Ok, K. M. (2023). Dalton Trans. 53, 105-114.]).

Isolated monomeric or dimeric cations of the type [SnIIF(LB0)3]+ or [SnIIF(LB0)2]+ with LB0 = neutral Lewis base have not yet been described in the literature, but a comparable mixed-valent tin(II)–tin(IV) compound also from SnF2 and 2,2′-bi­pyridine was previously described by Gurnani et al. (2013[Gurnani, C., Hector, A. L., Jager, E., Levason, W., Pugh, D. & Reid, G. (2013). Dalton Trans. 42, 8364-8374.]): [SnF(bipy)]2[SnF6]. In this compound, the first coordination sphere of the bivalent tin atom exhibits a fourfold, seesaw {SnN2F2} coordination with one additional, axial fluorine atom [d(Sn—F) = 2.510 (2) Å] instead of the axial water mol­ecule in the title compound and similar tin–fluorine [d(Sn—F)eq = 2.031 (2) Å] and tin–nitro­gen distances [d(Sn—N)eq = 2.298 (3) Å, d(Sn—N)ax = 2.335 (3) Å]. As in the title compound, dimerization proceeds via an asymmetrical 3c–4e bond [d(Sn⋯Feq) = 2.738 (3) Å], generating a similar, centrosymmetric four-membered Sn–F ring and trapezoid–pyramidally coordinated tin atoms. Regarding the centrosymmetric [SnF6]2− anion of this compound, the Sn—F distances are 1.956 (3) and 1.944 (3) Å for the terminal fluorine atoms and 1.995 (3) Å in case of the μ2-fluorine atom that adopts the axial position in the first coordination sphere of the bivalent tin atom.

5. Synthesis and crystallization

A mixture of approximately 157 mg (1 mmol) of SnF2 (Aldrich) and 156 mg (1 mmol) of 2,2′-bi­pyridine (Sigma Aldrich) was placed in a Petri dish and 5 ml of N,N-di­methyl­formamide (Sigma Aldrich) were added via a syringe. No elemental tin was added, nor was oxygen from the air excluded. The following processes were observed from day to day with a light microscope (Stemi 1000, Zeiss, Oberkochen, Germany). After two days small, colorless, block-like crystals of the title compound appeared, which, after two additional days, reached a size that was sufficient for a single crystal X-ray determination.

6. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. The positions of all H atoms were clearly identified in difference-Fourier maps. Those of the organic ligand were refined with calculated positions (C—H = 0.93 Å) and a common isotropic displacement parameter. The positions of the H atoms of the water mol­ecule were refined with a fixed O—H distance of 0.96 Å and a bond angle of 104.95° before they were fixed and allowed to ride on the parent O atom with an isotropic displacement parameter.

Table 2
Experimental details

Crystal data
Chemical formula [SnF(C10H8N2)(H2O)]2[SnF6]
Mr 856.47
Crystal system, space group Triclinic, P[\overline{1}]
Temperature (K) 100
a, b, c (Å) 7.4579 (4), 8.4601 (5), 9.8864 (6)
α, β, γ (°) 97.073 (2), 98.474 (2), 100.644 (2)
V3) 599.06 (6)
Z 1
Radiation type Mo Kα
μ (mm−1) 3.19
Crystal size (mm) 0.17 × 0.15 × 0.05
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.]
Tmin, Tmax 0.434, 0.713
No. of measured, independent and observed [I > 2σ(I)] reflections 28904, 2864, 2611
Rint 0.063
(sin θ/λ)max−1) 0.660
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.022, 0.053, 1.05
No. of reflections 2864
No. of parameters 172
H-atom treatment Only H-atom displacement parameters refined
Δρmax, Δρmin (e Å−3) 0.89, −0.69
Computer programs: APEX2 and SAINT (Bruker, 2009[Bruker (2009). APEX2 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2014/7 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), DIAMOND (Brandenburg, 2006[Brandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.]), Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

Bis[aqua(2,2'-bipyridine)fluoridotin(II)] hexafluoridotin(IV) top
Crystal data top
[SnF(C10H8N2)(H2O)]2[SnF6]Z = 1
Mr = 856.47F(000) = 406
Triclinic, P1Dx = 2.374 Mg m3
a = 7.4579 (4) ÅMo Kα radiation, λ = 0.71073 Å
b = 8.4601 (5) ÅCell parameters from 9888 reflections
c = 9.8864 (6) Åθ = 2.5–28.9°
α = 97.073 (2)°µ = 3.19 mm1
β = 98.474 (2)°T = 100 K
γ = 100.644 (2)°Plate, colourless
V = 599.06 (6) Å30.17 × 0.15 × 0.05 mm
Data collection top
Bruker APEXII CCD
diffractometer
2611 reflections with I > 2σ(I)
φ and ω scansRint = 0.063
Absorption correction: multi-scan
(SADABS; Krause et al., 2015
θmax = 28.0°, θmin = 3.0°
Tmin = 0.434, Tmax = 0.713h = 99
28904 measured reflectionsk = 1011
2864 independent reflectionsl = 1313
Refinement top
Refinement on F2Hydrogen site location: mixed
Least-squares matrix: fullOnly H-atom displacement parameters refined
R[F2 > 2σ(F2)] = 0.022 w = 1/[σ2(Fo2) + (0.024P)2 + 0.4078P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.053(Δ/σ)max < 0.001
S = 1.05Δρmax = 0.89 e Å3
2864 reflectionsΔρmin = 0.69 e Å3
172 parametersExtinction correction: SHELXL2014/7 (Sheldrick, 2015), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
0 restraintsExtinction coefficient: 0.0061 (6)
Primary atom site location: structure-invariant direct methods
Special details top

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

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Sn10.50001.00001.00000.01059 (8)
F110.22638 (19)0.91895 (18)0.96683 (15)0.0190 (3)
F120.5060 (2)0.96203 (18)0.80238 (14)0.0192 (3)
F130.4618 (2)1.21909 (17)0.98795 (15)0.0198 (3)
Sn20.21949 (2)0.11033 (2)0.42958 (2)0.01159 (8)
F210.05571 (18)0.10252 (17)0.41199 (14)0.0143 (3)
N10.2719 (3)0.3727 (2)0.3702 (2)0.0124 (4)
C10.3234 (3)0.3987 (3)0.2477 (3)0.0150 (5)
H10.33620.30830.18540.015 (2)*
C20.3578 (3)0.5513 (3)0.2103 (2)0.0161 (6)
H20.39390.56600.12370.015 (2)*
C30.3388 (3)0.6832 (3)0.3011 (3)0.0169 (5)
H30.35970.78950.27690.015 (2)*
C40.2890 (3)0.6585 (3)0.4277 (3)0.0149 (5)
H40.27610.74770.49140.015 (2)*
C50.2581 (3)0.5018 (3)0.4601 (2)0.0125 (5)
C60.2159 (3)0.4654 (3)0.5971 (2)0.0120 (5)
N20.2082 (3)0.3096 (2)0.6171 (2)0.0118 (4)
C70.1811 (3)0.2676 (3)0.7394 (2)0.0152 (5)
H70.17910.15820.75340.015 (2)*
C80.1557 (4)0.3780 (3)0.8464 (3)0.0174 (5)
H80.13430.34470.93190.015 (2)*
C90.1622 (3)0.5373 (3)0.8268 (3)0.0167 (5)
H90.14500.61510.89910.015 (2)*
C100.1940 (3)0.5838 (3)0.7003 (2)0.0135 (5)
H100.20060.69340.68520.015 (2)*
O10.0938 (2)0.0501 (2)0.18594 (17)0.0187 (4)
H1010.02290.06700.14120.073 (10)*
H1020.14810.00340.11260.073 (10)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Sn10.00996 (13)0.01193 (14)0.01123 (13)0.00387 (9)0.00360 (9)0.00239 (9)
F110.0116 (7)0.0237 (8)0.0199 (8)0.0027 (6)0.0030 (6)0.0024 (6)
F120.0216 (8)0.0231 (8)0.0133 (7)0.0042 (6)0.0053 (6)0.0027 (6)
F130.0255 (8)0.0155 (8)0.0224 (8)0.0096 (6)0.0069 (6)0.0062 (6)
Sn20.01071 (11)0.01044 (11)0.01397 (12)0.00390 (7)0.00208 (7)0.00069 (7)
F210.0110 (7)0.0161 (7)0.0165 (7)0.0044 (5)0.0035 (5)0.0015 (6)
N10.0104 (10)0.0133 (10)0.0134 (10)0.0023 (8)0.0027 (8)0.0004 (8)
C10.0130 (12)0.0165 (13)0.0145 (13)0.0023 (10)0.0017 (10)0.0009 (10)
C20.0120 (13)0.0223 (15)0.0150 (14)0.0043 (11)0.0022 (11)0.0054 (12)
C30.0146 (12)0.0152 (13)0.0214 (14)0.0020 (10)0.0025 (10)0.0071 (11)
C40.0138 (12)0.0124 (12)0.0186 (13)0.0050 (9)0.0019 (10)0.0003 (10)
C50.0066 (11)0.0152 (12)0.0161 (12)0.0031 (9)0.0020 (9)0.0023 (10)
C60.0091 (11)0.0115 (12)0.0145 (12)0.0020 (9)0.0004 (9)0.0009 (10)
N20.0128 (10)0.0107 (10)0.0111 (10)0.0024 (8)0.0003 (8)0.0010 (8)
C70.0156 (12)0.0148 (12)0.0132 (12)0.0008 (10)0.0010 (10)0.0024 (10)
C80.0168 (13)0.0225 (14)0.0118 (12)0.0012 (10)0.0026 (10)0.0027 (11)
C90.0156 (12)0.0190 (13)0.0141 (13)0.0044 (10)0.0014 (10)0.0027 (10)
C100.0136 (12)0.0121 (12)0.0142 (12)0.0032 (9)0.0009 (9)0.0000 (10)
O10.0162 (9)0.0269 (10)0.0132 (9)0.0068 (8)0.0037 (7)0.0016 (8)
Geometric parameters (Å, º) top
Sn1—F131.9418 (14)C3—H30.9500
Sn1—F13i1.9418 (14)C4—C51.389 (3)
Sn1—F12i1.9491 (14)C4—H40.9500
Sn1—F121.9491 (14)C5—C61.490 (3)
Sn1—F111.9931 (14)C6—N21.349 (3)
Sn1—F11i1.9931 (14)C6—C101.389 (3)
Sn2—F212.0221 (13)N2—C71.335 (3)
Sn2—N12.344 (2)C7—C81.381 (4)
Sn2—N22.369 (2)C7—H70.9500
Sn2—O12.4076 (17)C8—C91.378 (4)
N1—C51.350 (3)C8—H80.9500
N1—C11.354 (3)C9—C101.396 (3)
C1—C21.375 (3)C9—H90.9500
C1—H10.9500C10—H100.9500
C2—C31.386 (3)O1—H1010.9600
C2—H20.9500O1—H1020.9600
C3—C41.385 (4)
F13—Sn1—F13i180.0C4—C3—C2119.3 (2)
F13—Sn1—F12i89.28 (6)C4—C3—H3120.3
F13i—Sn1—F12i90.72 (6)C2—C3—H3120.3
F13—Sn1—F1290.72 (6)C3—C4—C5119.1 (2)
F13i—Sn1—F1289.28 (6)C3—C4—H4120.4
F12i—Sn1—F12180.00 (9)C5—C4—H4120.4
F13—Sn1—F1189.59 (6)N1—C5—C4121.6 (2)
F13i—Sn1—F1190.41 (6)N1—C5—C6115.7 (2)
F12i—Sn1—F1188.39 (6)C4—C5—C6122.7 (2)
F12—Sn1—F1191.61 (6)N2—C6—C10121.7 (2)
F13—Sn1—F11i90.41 (6)N2—C6—C5115.3 (2)
F13i—Sn1—F11i89.59 (6)C10—C6—C5122.9 (2)
F12i—Sn1—F11i91.62 (6)C7—N2—C6119.5 (2)
F12—Sn1—F11i88.39 (6)C7—N2—Sn2120.78 (16)
F11—Sn1—F11i180.00 (3)C6—N2—Sn2119.55 (15)
F21—Sn2—N191.52 (6)N2—C7—C8122.1 (2)
F21—Sn2—N278.11 (6)N2—C7—H7118.9
N1—Sn2—N268.50 (7)C8—C7—H7118.9
F21—Sn2—O173.42 (6)C9—C8—C7118.8 (2)
N1—Sn2—O179.62 (6)C9—C8—H8120.6
N2—Sn2—O1136.22 (6)C7—C8—H8120.6
C5—N1—C1118.7 (2)C8—C9—C10119.7 (2)
C5—N1—Sn2120.40 (15)C8—C9—H9120.2
C1—N1—Sn2120.81 (16)C10—C9—H9120.2
N1—C1—C2122.4 (2)C6—C10—C9118.1 (2)
N1—C1—H1118.8C6—C10—H10121.0
C2—C1—H1118.8C9—C10—H10121.0
C1—C2—C3118.8 (2)Sn2—O1—H101127.2
C1—C2—H2120.6Sn2—O1—H102127.9
C3—C2—H2120.6H101—O1—H102105.0
Symmetry code: (i) x+1, y+2, z+2.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O1—H101···F11ii0.961.752.705 (2)170
O1—H102···F11iii0.961.752.710 (2)174
Symmetry codes: (ii) x, y+1, z+1; (iii) x, y1, z1.
 

Acknowledgements

The Deutsche Forschungsgemeinschaft and the Government of Lower-Saxony are thanked for the funding of the diffractometer.

References

First citationBenner, G. & Hoppe, R. (1990). J. Fluor. Chem. 48, 219–227.  CrossRef ICSD CAS Web of Science Google Scholar
First citationBouacida, S., Merazig, H., Beghidja, A. & Beghidja, C. (2005). Acta Cryst. E61, m577–m579.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationBrandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBruker (2009). APEX2 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationCortijo, M., Bulicanu, V., Pedersen, K. S., Rouzières, M., Bendix, J., Clérac, R. & Hillard, E. (2017). Eur. J. Inorg. Chem. pp. 320–325.  Google Scholar
First citationGurnani, C., Hector, A. L., Jager, E., Levason, W., Pugh, D. & Reid, G. (2013). Dalton Trans. 42, 8364–8374.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationJung, A., Li, Y. & Ok, K. M. (2023). Dalton Trans. 53, 105–114.  Web of Science CSD CrossRef PubMed Google Scholar
First citationKleeberg, F. M., Zimmermann, L. W. & Schleid, T. (2022). J. Clust Sci. 33, 2489–2497.  Web of Science CSD CrossRef CAS Google Scholar
First citationKokunov, Yu. V., Gorbunova, Yu. E. & Razgonyaeva, G. A. (2007). Russ. J. Inorg. Chem. 52, 158–164.  Web of Science CrossRef Google Scholar
First citationKrause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.  Web of Science CSD CrossRef ICSD CAS IUCr Journals Google Scholar
First citationLermontov, S. A., Malkova, A. N., Lermontova, E. Kh. & Churakov, A. V. (2010). Russ. J. Gen. Chem. 80, 1825–1830.  Web of Science CrossRef CAS Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMusher, J. I. (1969). Angew. Chem. Int. Ed. Engl. 8, 54–68.  CrossRef CAS Web of Science Google Scholar
First citationPimentel, G. C. (1951). J. Chem. Phys. 19, 446–448.  CrossRef CAS Web of Science Google Scholar
First citationRundle, R. E. (1949). J. Chem. Phys. 17, 671–675.  CrossRef CAS Web of Science Google Scholar
First citationSheldrick, G. M. (2008). Acta Cryst. A64, 112–122.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationShimoni-Livny, L., Glusker, J. P. & Bock, C. W. (1998). Inorg. Chem. 37, 1853–1867.  Web of Science CrossRef CAS Google Scholar
First citationTaha, A., Liautard, B. & Granier, W. (1992). Acta Cryst. C48, 1929–1931.  CSD CrossRef CAS IUCr Journals Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar

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

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