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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Synthesis, crystal structure and Hirshfeld surface of bis­­(2-amino­pyridinium) hexa­chlorido­stannate(IV)

CROSSMARK_Color_square_no_text.svg

aEnvironmental Molecular, and Structural Chemistry Research Unit, University of, Constantine-1, 25000, Constantine, Algeria, and bLaboratory of Solid State Chemistry and Mossbauer Spectroscopy, Departement of, Chemistry and Biochemistry, Concordia University, 7141 Sherbrooke St. West, Montreal, H4B 1R6, QC, Canada
*Correspondence e-mail: rochdi.ghallab@gmail.com

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 1 June 2020; accepted 9 July 2020; online 17 July 2020)

In the title mol­ecular salt, (C5H7N2)2[SnCl6], the cation is protonated at the pyridine N atom and the complete dianion is generated by a crystallographic centre of symmetry. In the crystal, N—H⋯Cl hydrogen bonds link the components into a three-dimensional network built up from the stacking of alternate cationic and anionic layers. The nature of the inter­molecular inter­actions has been analysed in terms of the Hirshfeld surfaces of the cations and the anions. The thermal behaviour and the Raman spectrum of the title compound are reported.

1. Chemical context

So-called `zero-dimensional' hybrid perovskites are characterized by a structure formed by isolated inorganic octa­hedra (or bi­octa­hedra) and an organic cation (Cheng & Lin, 2010[Cheng, Z. & Lin, J. (2010). CrystEngComm, 12, 2646-2662.]). They are easy to prepare through simple techniques (Mitzi, 2004[Mitzi, D. B. (2004). J. Mater. Chem. 14, 2355-2365.]) and they combine the properties of the various organic and inorganic compounds, i.e. the flexibility of the organic part, and the thermal stability and the rigidity of the inorganic part, in a single material, by cooperative effects, to obtain properties that are more than just the sum of the initial properties: an organic/inorganic `synergy' is created. For example, in these hybrid materials, the organic part can have non-linear optical properties (Bi et al., 2008[Bi, W., Louvain, N., Mercier, N., Luc, J., Rau, I., Kajzar, F. & Sahraoui, B. (2008). Adv. Mater. 20, 1013-1017.]). Most of the physical properties come from the inorganic part, such as the electronic transport properties, the optical photoluminescence properties (Yangui et al., 2019[Yangui, A., Roccanova, R., Wu, Y., Du, M.-H. & Saparov, B. (2019). J. Phys. Chem. C, 123, 22470-22477.]), or even magnetic properties (Manser et al., 2016[Manser, J. S., Christians, J. A. & Kamat, P. V. (2016). Chem. Rev. 116, 12956-13008.]). As part of our studies in this area, we now describe the synthesis and structure of the title mol­ecular salt, (I)[link].

[Scheme 1]

2. Structural commentary

Compound (I)[link] with formula (C5H7N2)+2·[SnCl6]2−, crystallizes in the triclinic space group P[\overline{1}] (Fig. 1[link]).

[Figure 1]
Figure 1
The mol­ecular structure of the title compound with displacement ellipsoids drawn at the 30% probability level [Symmetry code: (i) −x, −y, −z].

In the synthesis, the oxidation number of tin changes from +2 to +4 such that the resultant tin(IV) atom is hexa­coordinated by chlorine atoms, generating a weakly distorted octa­hedron in which the metal ion lies on a crystallographic inversion centre: the length of the Sn—Cl bonds varies from 2.4216 (4) to 2.4474 (5) Å. As for the Cl—Sn—Cl angles, the discrepancy of about ± 1° [89.109 (18)–90.805 (16)°] compared to the 90° value angle of a regular octa­hedron shows that the angular distortion is very small. These values are comparable to those of the same anion associated with other types of cations (BelhajSalah et al., 2018[BelhajSalah, S., Abdelbaky, M. S. M., García-Granda, S., Essalah, K., Ben Nasr, C. & Mrad, M. L. (2018). Solid State Sci. 86, 77-85.]). The absence of larger distortions can probably be attributed to the fact that the hexa­chloro­stannate(IV) anions are free, i.e. none of the chloride ions are bridging although they do accept N—H⋯Cl hydrogen bonds from the organic cations, which ensures charge balance.

In the pyridinium ring of the cation, the C—C bond lengths vary from 1.328 (3) to 1.405 (3) Å and the C—N bond lengths are 1.341 (3) Å and 1.344 (2) Å. The values of the C—C—C angles in the pyridinium ring vary from 118.9 (2) to 120.9 (2)° whereas the C—N—C angle is 124.30 (18)°: the larger angle can be attributed to the protonation of the N atom. These values are comparable with those of the same cation associated with other types of anions (Rao et al., 2011[Rao, A. S., Baruah, U. & Das, S. K. (2011). Inorg. Chim. Acta, 372, 206-212.]).

3. Supra­molecular features

The special position of tin(IV) in the crystal of (I)[link] gives rise to an alternation of cationic and anionic layers lying parallel to the (001) plane (Fig. 2[link]a). The inter­molecular inter­actions in (I)[link] were analysed using PLATON (Spek, 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.]), which shows that the structural cohesion in the crystalline structure of the compound (I)[link] is ensured by N—H⋯Cl hydrogen bonds: Fig. 2[link]a. The distances and the angles describing these inter­actions are presented in Table 1[link].

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1⋯Cl1i 0.77 (2) 2.79 (3) 3.3155 (18) 128 (3)
N2—H2A⋯Cl1ii 0.86 2.62 3.406 (2) 152
N2—H2B⋯Cl2iii 0.86 2.52 3.358 (2) 166
Symmetry codes: (i) x, y+1, z; (ii) -x-1, -y+1, -z; (iii) -x-1, -y+1, -z+1.
[Figure 2]
Figure 2
(a) Hydrogen bonds [symmetry codes: (i) x, y + 1, z; (ii) −x − 1, −y + 1, −z; (iii) −x − 1, −y + 1, −z + 1] (blue dashed lines), π-stacking inter­actions (symmetry operation: −1 − x, 1 − y, 1 − z) and Y—XCg (symmetry operation: x, −1 + y, z) (red dashed lines) in the unit cell of compound (I)[link]. (b) A view of the ring motifs along the c axis with the strongest hydrogen bonds and π-stacking inter­actions indicated by red dotted lines.

The combination of N2—H2A⋯Cl1 and N2—H2B⋯Cl2 hydrogen bonds generates a chain of rings propagating along the [001] direction with a graph-set pattern of R44(12) (Etter et al., 1990[Etter, M. C., MacDonald, J. C. & Bernstein, J. (1990). Acta Cryst. B46, 256-262.]), Fig. 2[link](b). The cohesion between chains is ensured by π-stacking inter­actions between centrosymmetrically related aromatic rings of the cations: Cg1⋯Cg1 = 3.552 (13) Å; inter-planar angle α = 0.0 (11)°; slippage = 1.246 Å. We also note the presence of a YXCg1 type inter­action between Sn1—Cl2 and Cg1 at an XCg distance of 3.6581 (11) Å [Fig. 2[link](a)].

4. Hirshfeld surface analysis:

To further characterize the inter­molecular inter­actions in (I)[link], the Hirshfeld surface method was used (Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]). In addition, its two-dimensional fingerprints (Spackman & McKinnon, 2002[Spackman, M. A. & McKinnon, J. J. (2002). CrystEngComm, 4, 378-392.]) were calculated using the program Crystal Explorer 17 (Turner et al., 2017[Turner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.]). The dnorm representation mode was used in which red spots identify close contacts; in the white areas, the distance separating the neighboring atoms approaches the sum of the van der Waals radii of the concerned atoms whereas blue areas illustrate areas where neighbouring atoms are too far apart to inter­act significantly with each other. The presence of the adjacent red and blue triangles, obtained by using the shape index as a representation mode, demonstrates the presence of ππ and YXπ type inter­actions.

The Hirshfeld surface [Fig. 3[link](a)] shows red spots corresponding to H⋯Cl/Cl⋯H close contacts, which are due to the N—H⋯Cl hydrogen bonds. The presence of the adjacent red and blue triangles in Fig. 3[link](b) demonstrates the presence of the Cg1⋯Cg1 and Sn—Cl2⋯Cg1 inter­actions. The contribution of different kinds of inter­atomic contacts to the Hirshfeld surfaces of the individual cations and anions is shown in the fingerprint plots in Fig. 4[link] and Fig. 5[link], respectively. These inter­actions are ensured by 47.3% of hydrogen bonds (H⋯Cl), 3.2% of YX⋯ type (N⋯Cl and C⋯Cl), 6.6% of ππ stacking type (C⋯C and C⋯N/N⋯C), 15.6% of C—H⋯π type (C⋯H/H⋯C), 6.2% of N—H⋯π type (N⋯H/H⋯N) and 21.1% of H⋯H van der Waals inter­actions. The two-dimensional fingerprint analysis for the anionic moieties reveals that hydrogen bonds (Cl⋯H) represent 93.8%, YXπ type inter­actions represent 4.4% (Cl⋯N and Cl⋯C) and van der Waals inter­actions of the Cl⋯Cl type represent 1.8% of the surface contacts.

[Figure 3]
Figure 3
(a) A view of the Hirshfeld surface mapped over dnorm for compound (I)[link] in the range −0.3387 to +1.0913 arbitrary units, highlighting the N—H⋯Cl inter­actions and (b) the Hirshfeld surface of the cation mapped over shape-index in the range −1.00 to +1.00 arbitrary units.
[Figure 4]
Figure 4
Two-dimensional fingerprint plots for the cation of the title compound, and delineated into the principal contributions of H⋯Cl, C⋯C, C⋯N/N⋯C, N⋯Cl and C⋯Cl contacts. Other significant contacts are H⋯H (21.1%), H⋯C/C⋯H (15.6%) and H⋯N/N⋯H (6.2%).
[Figure 5]
Figure 5
Two-dimensional fingerprint plots for the anion of the title compound, and those delineated into the principal contributions of Cl⋯H, Cl⋯C, Cl⋯N and Cl⋯Cl contacts.

5. Database survey

A search of the Cambridge Structural Database (CSD Version 5.41; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) for structures similar to (I)[link] gave several compounds such as 2-amino­pyridinium hexa­chloro­bis­muth(III) (Rao et al., 2011[Rao, A. S., Baruah, U. & Das, S. K. (2011). Inorg. Chim. Acta, 372, 206-212.]), 2-amino­pyridinium hexa­chloro­indium(III) (Jin et al., 2011[Jin, X.-D., Sun, L.-C., Wang, H.-B. & Ge, C.-H. (2011). Acta Cryst. E67, m1602.]), 4-amino­pyridinium hexa­chloro­anti­monate(V) (Kulicka et al., 2006[Kulicka, B., Jakubas, R., Pietraszko, A., Medycki, W. & Świergiel, J. (2006). J. Mol. Struct. 783, 88-95.]) and 4-amino­pyridinium hexa­chloro­stannate(IV) (Rademeyer et al., 2007[Rademeyer, M., Lemmerer, A. & Billing, D. G. (2007). Acta Cryst. C63, m289-m292.]) among others, but the last of these (refcode RIGDER) is of particular inter­est: RIGDER and (I)[link] both crystallize in space group P[\overline{1}] where the [SnCl6]2− anions are associated with special positions and an organic–inorganic layered structure lying parallel to the (001) plane results.

Crystalline cohesion in RIGDER and (I)[link] is ensured by dipole–dipole inter­actions and hydrogen bonds of the N—H⋯Cl type with a slight difference in the donor–acceptor angles and distances of the two compounds. The different arrangement of the nitro­gen atoms in the cation in RIGDER leads to much weaker ππ stacking compared to (I)[link]: the centroid separations are 4.24 (1) and 3.552 (13) Å, respectively. We also notice a slight difference between the two compounds in the inter­action percentages calculated by the Hirshfeld surface analysis (see Table S1 in the supporting information).

6. Thermal analysis

In order to investigate the thermal stability of (I)[link], thermogravimetric analysis (DTA/TGA) was performed under an N2 atmosphere at a heating rate of 10°C min−1 in the temperature range from 25 to 500°C. The thermogram of (I)[link] (see Fig. S2 in the supporting information) shows that the compound loses 64.4% of its mass in the temperature range of 270–304°C. The mass loss can be attributed to the degradation of the organic entity and two chlorine atoms (Janiak & Blazejowski, 1990[Janiak, T. & Blazejowski, J. (1990). Thermochim. Acta, 157, 137-154.]) to leave a reside of SnCl4.

7. Synthesis and crystallization

Tin(II) chloride dihydrate (2.25 mmol) was mixed with 2-amino­pyridine (0.94 mmol) and a few drops of hydro­chloric acid in an aliquot of distilled water in 1:1 molar ratio was added. After stirring, the mixture was poured into a vial (biotage microwave vial 2–5 ml) that was put in an oven for three days at 393 K. Upon cooling, prism-shaped crystals of (I)[link] were obtained and separated using an optical microscope. 1H NMR (δ ppm), 400 MHz, CDCl3): 8.16 (br s, 2H, NH2), 7.95–7.89 (m, 2H CH Py), 7.03 (d, JHH = 8.9 Hz,1H CH Py), 6.84 (t, JHH = 6.6 Hz, 1H CH Py). 13C NMR (δ ppm), 125 MHz, CDCl3): 154.6 (quat C Py), 144.4 (CH Py), 136.1 (CH Py), 113.8 (CH Py), 112.5 (CH Py).

The raman spectrum for (I)[link] (Fig. 6[link]) was recorded in the frequency range 4000–60 cm−1. The Py, ν, δ, γ and τ are: pyridine ring, stretching, in-plane bending, out-of-plane bending and torsion, respectively. RS (cm−1): 3334 ν(N—H), 3215 ν (N—H), 3106 ν(C—H), 1657 ν(py)+δ(N—H)+δ(NH2), 1620 ν(py), 1542 ν(py), 1472 ν(py)+δ(C—H), 1412 ν(py)+δ(C—H), 1378 ν(py)+δ(C—H), 1324 ν(py)+δ(C—H), 1239 ν(py)+δ(C—H), 1164 δ(py)+δ(C—H), 1120 δ(py)+δ(C—H), 996 δ(py), 846 Pyridine ring breathing+γ(C—H), 623 γ(py), 551 γ(py), 406 γ(py), 384 γ(py), 305 ν1 (Sn—Cl), 216 ν2 (Sn—Cl), 106 τ(py)+ ν(N—H⋯Cl) (Shaw et al., 1988[Shaw, R. A., Castro, C., Dutler, R., Rauk, A. & Wieser, H. (1988). J. Chem. Phys. 89, 716-731.]; Ureña et al., 2003[Ureña, F. P., Gómez, M. F., González, J. J. L. & Torres, E. M. (2003). Spectrochim. Acta A Mol. Biomol. Spectrosc. 59, 2815-2839.]; Cook, 1961[Cook, D. (1961). Can. J. Chem. 39, 2009-2024.]).

[Figure 6]
Figure 6
Raman spectrum of compound (I)[link].

8. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. The C-bound H atoms and the anine H atom were placed geometrically and refined as riding atoms [C—H = 0.93 Å and Uiso(H) = 1.2Ueq(C)]; the pyridine N—H atom was located in a difference map and its position was freely refined.

Table 2
Experimental details

Crystal data
Chemical formula (C5H7N2)2[SnCl6]
Mr 521.64
Crystal system, space group Triclinic, P[\overline{1}]
Temperature (K) 298
a, b, c (Å) 7.4537 (1), 8.0674 (1), 8.1025 (1)
α, β, γ (°) 83.791 (1), 82.591 (1), 71.407 (1)
V3) 456.77 (1)
Z 1
Radiation type Mo Kα
μ (mm−1) 2.27
Crystal size (mm) 0.08 × 0.08 × 0.07
 
Data collection
Diffractometer Bruker SMART APEXII area detector
Absorption correction Multi-scan (SADABS; Bruker, 2016[Bruker (2016). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.834, 0.853
No. of measured, independent and observed [I > 2σ(I)] reflections 10455, 2011, 1910
Rint 0.021
(sin θ/λ)max−1) 0.641
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.017, 0.037, 1.06
No. of reflections 2011
No. of parameters 101
No. of restraints 36
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.56, −0.34
Computer programs: APEX2 and SAINT (Bruker, 2016[Bruker (2016). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SIR2004 (Burla et al., 2007[Burla, M. C., Caliandro, R., Camalli, M., Carrozzini, B., Cascarano, G. L., De Caro, L., Giacovazzo, C., Polidori, G., Siliqi, D. & Spagna, R. (2007). J. Appl. Cryst. 40, 609-613.]), SHELXL (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]) and OLEX2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]).

Supporting information


Computing details top

Data collection: APEX2 (Bruker, 2016); cell refinement: SAINT (Bruker, 2016); data reduction: SAINT (Bruker, 2016); program(s) used to solve structure: SIR2004 (Burla et al., 2007); program(s) used to refine structure: SHELXL (Sheldrick, 2015); molecular graphics: OLEX2 (Dolomanov et al., 2009); software used to prepare material for publication: OLEX2 (Dolomanov et al., 2009).

Bis(2-aminopyridinium) hexachloridostannate(IV) top
Crystal data top
2C5H7N2+·Cl6Sn2Z = 1
Mr = 521.64F(000) = 254
Triclinic, P1Dx = 1.896 Mg m3
a = 7.4537 (1) ÅMo Kα radiation, λ = 0.71073 Å
b = 8.0674 (1) ÅCell parameters from 1910 reflections
c = 8.1025 (1) Åθ = 3.6–27.1°
α = 83.791 (1)°µ = 2.27 mm1
β = 82.591 (1)°T = 298 K
γ = 71.407 (1)°Prism, colourless
V = 456.77 (1) Å30.08 × 0.08 × 0.07 mm
Data collection top
Bruker SMART APEXII area detector
diffractometer
2011 independent reflections
Radiation source: microfocus sealed X-ray tube, Incoatec Iµs1910 reflections with I > 2σ(I)
Mirror optics monochromatorRint = 0.021
Detector resolution: 7.9 pixels mm-1θmax = 27.1°, θmin = 3.6°
ω and φ scansh = 99
Absorption correction: multi-scan
(SADABS; Bruker, 2016)
k = 1010
Tmin = 0.834, Tmax = 0.853l = 1010
10455 measured reflections
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullHydrogen site location: mixed
R[F2 > 2σ(F2)] = 0.017H atoms treated by a mixture of independent and constrained refinement
wR(F2) = 0.037 w = 1/[σ2(Fo2) + (0.0121P)2 + 0.2078P]
where P = (Fo2 + 2Fc2)/3
S = 1.06(Δ/σ)max < 0.001
2011 reflectionsΔρmax = 0.56 e Å3
101 parametersΔρmin = 0.33 e Å3
36 restraints
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
Sn10.0000000.0000000.0000000.03137 (6)
Cl30.29282 (7)0.24458 (7)0.00146 (6)0.04838 (12)
Cl10.17021 (7)0.18776 (7)0.08097 (6)0.04442 (11)
Cl20.07788 (7)0.07356 (7)0.29243 (5)0.04473 (12)
C10.4688 (3)0.7533 (2)0.4088 (2)0.0364 (4)
C20.4174 (3)0.6744 (3)0.5662 (2)0.0455 (5)
H20.4813580.7262820.6630520.055*
C30.2718 (3)0.5200 (3)0.5747 (3)0.0538 (6)
H30.2377500.4659020.6783770.065*
C40.1736 (3)0.4425 (3)0.4298 (3)0.0540 (5)
H40.0754940.3370510.4361490.065*
C50.2228 (3)0.5220 (3)0.2827 (3)0.0498 (5)
H50.1571760.4733790.1849450.060*
N10.3668 (3)0.6723 (2)0.2748 (2)0.0414 (4)
N20.6083 (3)0.9016 (2)0.3872 (3)0.0568 (5)
H2A0.6331480.9454580.2880640.068*
H2B0.6741390.9542020.4723300.068*
H10.397 (4)0.716 (3)0.189 (3)0.061 (8)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Sn10.02900 (9)0.04010 (10)0.02432 (9)0.01029 (7)0.00234 (6)0.00090 (6)
Cl30.0391 (2)0.0533 (3)0.0432 (3)0.0020 (2)0.0084 (2)0.0073 (2)
Cl10.0495 (3)0.0531 (3)0.0382 (2)0.0260 (2)0.0058 (2)0.0026 (2)
Cl20.0467 (3)0.0600 (3)0.0261 (2)0.0177 (2)0.00014 (18)0.00261 (19)
C10.0335 (9)0.0424 (10)0.0374 (9)0.0171 (8)0.0039 (7)0.0036 (8)
C20.0521 (12)0.0664 (13)0.0283 (9)0.0338 (11)0.0015 (8)0.0027 (8)
C30.0577 (13)0.0643 (14)0.0501 (12)0.0345 (12)0.0251 (10)0.0215 (10)
C40.0444 (12)0.0477 (12)0.0717 (15)0.0166 (10)0.0112 (10)0.0014 (10)
C50.0458 (11)0.0468 (12)0.0570 (13)0.0158 (10)0.0039 (10)0.0114 (10)
N10.0500 (10)0.0482 (10)0.0288 (8)0.0199 (8)0.0027 (7)0.0015 (7)
N20.0529 (11)0.0540 (11)0.0573 (11)0.0035 (9)0.0117 (9)0.0105 (9)
Geometric parameters (Å, º) top
Sn1—Cl3i2.4315 (5)C2—C31.369 (3)
Sn1—Cl32.4315 (5)C3—H30.9300
Sn1—Cl12.4474 (4)C3—C41.397 (3)
Sn1—Cl1i2.4474 (4)C4—H40.9300
Sn1—Cl2i2.4216 (4)C4—C51.328 (3)
Sn1—Cl22.4216 (4)C5—H50.9300
C1—C21.405 (3)C5—N11.341 (3)
C1—N11.344 (2)N1—H10.77 (2)
C1—N21.324 (3)N2—H2A0.8600
C2—H20.9300N2—H2B0.8600
Cl3i—Sn1—Cl3180.0C1—C2—H2120.6
Cl3i—Sn1—Cl190.889 (18)C3—C2—C1118.89 (19)
Cl3—Sn1—Cl1i90.891 (18)C3—C2—H2120.6
Cl3—Sn1—Cl189.109 (18)C2—C3—H3119.6
Cl3i—Sn1—Cl1i89.111 (18)C2—C3—C4120.85 (19)
Cl1i—Sn1—Cl1180.0C4—C3—H3119.6
Cl2—Sn1—Cl3i89.575 (17)C3—C4—H4120.6
Cl2i—Sn1—Cl3i90.425 (17)C5—C4—C3118.9 (2)
Cl2i—Sn1—Cl389.575 (17)C5—C4—H4120.6
Cl2—Sn1—Cl390.425 (17)C4—C5—H5119.9
Cl2—Sn1—Cl190.805 (16)C4—C5—N1120.1 (2)
Cl2i—Sn1—Cl1i90.804 (16)N1—C5—H5119.9
Cl2i—Sn1—Cl189.196 (16)C1—N1—H1116.4 (19)
Cl2—Sn1—Cl1i89.195 (16)C5—N1—C1124.30 (18)
Cl2i—Sn1—Cl2180.0C5—N1—H1119.3 (19)
N1—C1—C2116.94 (18)C1—N2—H2A120.0
N2—C1—C2123.55 (18)C1—N2—H2B120.0
N2—C1—N1119.50 (18)H2A—N2—H2B120.0
Symmetry code: (i) x, y, z.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N1—H1···Cl1ii0.77 (2)2.79 (3)3.3155 (18)128 (3)
N2—H2A···Cl1iii0.862.623.406 (2)152
N2—H2B···Cl2iv0.862.523.358 (2)166
Symmetry codes: (ii) x, y+1, z; (iii) x1, y+1, z; (iv) x1, y+1, z+1.
 

Acknowledgements

Thanks are due to Jean-Claude Daran, Eric Manoury, Eric Deydier and Gabor Molnar from LCC Toulouse France for their technical support.

Funding information

Funding for this research was provided by: Unité de Recherche de Chimie de l'Environnement, Moléculaire et Structurale (UR.CHEMS); Direction Générale de la Recherche Scientifique et du Développement Technologique.

References

First citationBelhajSalah, S., Abdelbaky, M. S. M., García-Granda, S., Essalah, K., Ben Nasr, C. & Mrad, M. L. (2018). Solid State Sci. 86, 77–85.  Web of Science CSD CrossRef CAS Google Scholar
First citationBi, W., Louvain, N., Mercier, N., Luc, J., Rau, I., Kajzar, F. & Sahraoui, B. (2008). Adv. Mater. 20, 1013–1017.  Web of Science CSD CrossRef CAS Google Scholar
First citationBruker (2016). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBurla, M. C., Caliandro, R., Camalli, M., Carrozzini, B., Cascarano, G. L., De Caro, L., Giacovazzo, C., Polidori, G., Siliqi, D. & Spagna, R. (2007). J. Appl. Cryst. 40, 609–613.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCheng, Z. & Lin, J. (2010). CrystEngComm, 12, 2646–2662.  Web of Science CrossRef CAS Google Scholar
First citationCook, D. (1961). Can. J. Chem. 39, 2009–2024.  CrossRef CAS Web of Science Google Scholar
First citationDolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339–341.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEtter, M. C., MacDonald, J. C. & Bernstein, J. (1990). Acta Cryst. B46, 256–262.  CrossRef ICSD CAS Web of Science IUCr Journals Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationJaniak, T. & Blazejowski, J. (1990). Thermochim. Acta, 157, 137–154.  CrossRef CAS Web of Science Google Scholar
First citationJin, X.-D., Sun, L.-C., Wang, H.-B. & Ge, C.-H. (2011). Acta Cryst. E67, m1602.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationKulicka, B., Jakubas, R., Pietraszko, A., Medycki, W. & Świergiel, J. (2006). J. Mol. Struct. 783, 88–95.  Web of Science CSD CrossRef CAS Google Scholar
First citationManser, J. S., Christians, J. A. & Kamat, P. V. (2016). Chem. Rev. 116, 12956–13008.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMitzi, D. B. (2004). J. Mater. Chem. 14, 2355–2365.  Web of Science CrossRef CAS Google Scholar
First citationRademeyer, M., Lemmerer, A. & Billing, D. G. (2007). Acta Cryst. C63, m289–m292.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationRao, A. S., Baruah, U. & Das, S. K. (2011). Inorg. Chim. Acta, 372, 206–212.  Web of Science CSD CrossRef CAS Google Scholar
First citationShaw, R. A., Castro, C., Dutler, R., Rauk, A. & Wieser, H. (1988). J. Chem. Phys. 89, 716–731.  CrossRef CAS Web of Science Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSpackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19–32.  Web of Science CrossRef CAS Google Scholar
First citationSpackman, M. A. & McKinnon, J. J. (2002). CrystEngComm, 4, 378–392.  Web of Science CrossRef CAS Google Scholar
First citationSpek, A. L. (2020). Acta Cryst. E76, 1–11.  Web of Science CrossRef IUCr Journals Google Scholar
First citationTurner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.  Google Scholar
First citationUreña, F. P., Gómez, M. F., González, J. J. L. & Torres, E. M. (2003). Spectrochim. Acta A Mol. Biomol. Spectrosc. 59, 2815–2839.  Web of Science PubMed Google Scholar
First citationYangui, A., Roccanova, R., Wu, Y., Du, M.-H. & Saparov, B. (2019). J. Phys. Chem. C, 123, 22470–22477.  Web of Science CSD CrossRef CAS 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