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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Bis(2-methyl­pyridinium) tetra­chlorido­cuprate(II): synthesis, structure and Hirshfeld surface analysis

crossmark logo

aDepartment of Chemistry, School of Sciences, Indrashil University, Rajpur, Gujarat, 382740, India
*Correspondence e-mail: rajeshbhosale24@gmail.com, j.prakashareddy@gmail.com

Edited by G. Diaz de Delgado, Universidad de Los Andes, Venezuela (Received 21 December 2020; accepted 17 June 2021; online 25 June 2021)

The title compound, (C6H8N)2[CuCl4], crystallizes in the monoclinic space group I2/c. The coordination around the copper atom is a distorted tetra­hedron. The 2-methyl­pyridinium ion (C6H8N+) inter­acts with the tetra­chloro­cuprate anion through N—H⋯Cl and C—H(phen­yl)⋯Cl contacts, forming a hydrogen-bonded layer-like structure. The supra­molecular structure is further stabilized by C—H(meth­yl)⋯Cl inter­actions between the layers.

1. Chemical context

Supra­molecular organic and inorganic chemistry have been studied both from the fundamental as well as the application point of view, which is evident from the literature (Ziach et al., 2018[Ziach, K., Chollet, C., Parissi, V., Prabhakaran, P., Marchivie, M., Corvaglia, V., Bose, P. P., Laxmi-Reddy, K., Godde, F., Schmitter, J.-M., Chaignepain, S., Pourquier, P. & Huc, I. (2018). Nat. Chem. 10, 511-518.]; Thorat et al., 2013[Thorat, V. H., Ingole, T. S., Vijayadas, K. N., Nair, R. V., Kale, S. S., Ramesh, V. V. E., Davis, H. C., Prabhakaran, P., Gonnade, R. G., Gawade, R. L., Puranik, V. G., Rajamohanan, P. R. & Sanjayan, G. J. (2013). Eur. J. Org. Chem. pp. 3529-3542.]; Burslem et al., 2016[Burslem, G. M., Kyle, H. F., Prabhakaran, P., Breeze, A. L., Edwards, T. A., Warriner, S. L., Nelson, A. & Wilson, A. J. (2016). Org. Biomol. Chem. 14, 3782-3786.]). With the surge in the number of compounds reported, potential applications of supra­molecular inorganic materials in energy storage, separation, catalysis, sensors, mol­ecular magnets, optoelectronic materials, etc., have attracted greater attention in recent years (Mueller et al., 2006[Mueller, U., Schubert, M., Teich, F., Puetter, H., Schierle-Arndt, K. & Pastré, J. (2006). J. Mater. Chem. 16, 626-636.]; Wan et al., 2006[Wan, Y., Yang, H. & Zhao, D. (2006). Acc. Chem. Res. 39, 423-432.]; Férey et al., 2003[Férey, G., Latroche, M., Serre, C., Millange, F., Loiseau, T. & Percheron-Guégan, A. (2003). Chem. Commun. pp. 2976-2977.]; James, 2003[James, S. L. (2003). Chem. Soc. Rev. 32, 276-288.]; Eddaoudi et al., 2002[Eddaoudi, M., Kim, J., Rosi, N., Vodak, D., Wachter, J., O'Keeffe, M. & Yaghi, O. M. (2002). Science, 295, 469-472.]; Ruben et al., 2005[Ruben, M., Ziener, U., Lehn, J. M., Ksenofontov, V., Gütlich, P. & Vaughan, G. B. M. (2005). Chem. Eur. J. 11, 94-100.], Kitagawa et al., 2004[Kitagawa, S., Kitaura, R. & Noro, S. I. (2004). Angew. Chem. Int. Ed. 43, 2334-2375.], Stavila et al., 2014[Stavila, V., Talin, A. A. & Allendorf, M. D. (2014). Chem. Soc. Rev. 43, 5994-6010.]). Because of the divergent combination of ligands and metal salts, an enormous number of structural architectures with different sizes and shapes could be constructed (Moulton & Zaworotko, 2001[Moulton, B. & Zaworotko, M. J. (2001). Chem. Rev. 101, 1629-1658.]). The special characteristics and features such as ease of synthesis of the material, geometrically well-defined structures, exceptional tunability, post-synthetic modification, along with robustness of the material resulting from strong directional bonding, produce new opportunities and offer a unique platform amenable to the synthesis of more and more functional solids. For example, Adams et al. (2005[Adams, C. J., Crawford, P. C., Orpen, A. G., Podesta, T. J. & Salt, B. (2005). Chem. Commun. pp. 2457-2458.]) reported the synthesis of coordination compounds using a new synthetic route involving a thermal de­hydro­chlorination reaction in crystals of a pyridinium chloro­metallate bicomponent system, i.e., anionic metal complexes and organic cations.

As part of ongoing studies in our group (PrakashaReddy & Pedireddi, 2007[PrakashaReddy, J. & Pedireddi, V. R. (2007). Eur. J. Inorg. Chem. pp. 1150-1158.]; Reddy et al., 2014[Reddy, J. P., Swain, D. & Pedireddi, V. R. (2014). Cryst. Growth Des. 14, 5064-5071.]), the synthesis of coord­ination complexes using pyridine ligands has been reported. Hence, we further extended our studies to utilize the pyrid­in­ium ligand and to study in situ the single-crystal-to-single-crystal transition (SCSCT) to investigate the reaction mechanism. In our endeavours to synthesize new functional solids, using a transition-metal anion and a pyridinium cation, we have chosen the CuCl2 and 2-methyl­pyridinium salt complex. Herein, we report the synthesis and crystal structure of a bis­(2-methyl­pyridinium) tetra­chloro­cuprate coordination complex.

[Scheme 1]

2. Structural commentary

The title complex crystallizes in the monoclinic space group I2/c. Since the Cu2+ cation occupies a special position, the asymmetric unit consists of a 2-methyl­pyridinium cation, [2-Me(Py)H]+, and half of a tetra­chloro­cuprate(II) anion, [CuCl4]2–. The mol­ecular structure of the complex along with the atom-labelling scheme is shown in Fig. 1[link]. Each copper center is four-coordinated by chlorine anions and adopts a distorted tetra­hedral geometry. Structural analysis shows that the Cl—Cu—Cl angles vary from 98.55 (2) to 137.4 (3)° with four angles smaller and two larger than the standard tetra­hedral angle. A plausible reason for a larger deviation from the standard 109.5° might be due to the N—H⋯Cl and C—H⋯Cl inter­actions. A similar marked deviation from the standard tetra­hedral angle has been previously observed by other research groups (Wyrzykowski et al., 2011[Wyrzykowski, D., Wera, M., Sikorski, A., Jacewicz, D. & Chmurzyński, L. (2011). Cent. Eur. J. Chem. 9, 1096-1101.]; Jasrotia et al., 2018[Jasrotia, D., Verma, S. K., Sridhar, B., Alvi, P. A. & Kumar, A. (2018). Mater. Chem. Phys. 207, 98-104.]). The Cu—Cl bond lengths [Cu1—Cl1 = 2.250 (1) Å, Cu1—Cl2 = 2.249 (1) Å] agree well with those reported for other structures (Marsh et al., 1982[Marsh, W. E., Hatfield, W. E. & Hodgson, D. J. (1982). Inorg. Chem. 21, 2679-2684.]; Dodds et al., 2018[Dodds, C. A. & Kennedy, A. R. (2018). Acta Cryst. E74, 1369-1372.]; Molano et al., 2020[Molano, M. F., Lorett Velasquez, V. P., Erben, M. F., Nossa González, D. L., Loaiza, A. E., Echeverría, G. A., Piro, O. E., Tobón, Y. A., Ben Tayeb, K. & Gómez Castaño, J. A. (2020). Acta Cryst. E76, 148-154.]; Reddy, 2020[Reddy, J. P. (2020). Acta Cryst. E76, 1771-1774.]). The intra­molecular Car—Car bond lengths in the [2-Me(Py)H]+ fall in the range 1.370 (3)–1.395 (3) Å. The N1—C2 and N1—C6 bond lengths are 1.350 (2) and 1.346 (2) Å, respectively.

[Figure 1]
Figure 1
The mol­ecular structure of the title compound, showing the atom labelling and displacement ellipsoids drawn at the 50% probability level. Hydrogen atoms are shown as small spheres of arbitrary size. [Symmetry code: (i) 1 − x, y, [{1\over 2}] − z.]

3. Supra­molecular features and Hirshfeld surface analysis

In the crystal, complex mol­ecules related by the twofold rotation axis are connected by pairs of N—H⋯Cl and Car—H⋯Cl inter­actions through a protonated N and an aromatic hydrogen attached to the carbon atom with the chloride ligand bonded to copper, forming a monomeric unit. These units inter­act with adjacent ones through Car—H⋯Cl hydrogen bonding (Table 1[link], Fig. 2[link]). The N—H⋯Cl and C—H⋯Cl distances and associated bond angles lie within the ranges observed for other similar inter­actions reported in the literature (Adams et al., 2005[Adams, C. J., Crawford, P. C., Orpen, A. G., Podesta, T. J. & Salt, B. (2005). Chem. Commun. pp. 2457-2458.]; Vittaya et al., 2015[Vittaya, L., Leesakul, N., Saithong, S. & Chainok, K. (2015). Acta Cryst. E71, m201-m202.]; Wyrzykowski et al., 2011[Wyrzykowski, D., Wera, M., Sikorski, A., Jacewicz, D. & Chmurzyński, L. (2011). Cent. Eur. J. Chem. 9, 1096-1101.]; Jasrotia et al., 2018[Jasrotia, D., Verma, S. K., Sridhar, B., Alvi, P. A. & Kumar, A. (2018). Mater. Chem. Phys. 207, 98-104.]). The supra­molecular structure is further stabilized by Cmeth­yl—H⋯Cl inter­actions involving hydrogens of the methyl group and chlorides bonded to copper, generating layers along the crystallographic b axis (Fig. 3[link]).

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1⋯Cl1i 0.88 2.93 3.4297 (16) 118
N1—H1⋯Cl2i 0.88 2.41 3.2050 (16) 150
C6—H6⋯Cl1ii 0.95 2.62 3.453 (2) 147
C7—H7A⋯Cl2 0.98 2.92 3.850 (2) 159
C7—H7B⋯Cl2iii 0.98 2.98 3.872 (2) 151
Symmetry codes: (i) [-x+1, -y, -z+1]; (ii) [x+{\script{1\over 2}}, y-{\script{1\over 2}}, z+{\script{1\over 2}}]; (iii) [-x+1, -y+1, -z+1].
[Figure 2]
Figure 2
The N—H⋯Cl and C—H⋯Cl inter­actions between cations and anions in the crystal structure of the title compound.
[Figure 3]
Figure 3
The crystal packing of the title compound viewed along the b axis with inter­molecular contacts shown as dashed lines.

To further investigate the inter­molecular inter­actions present in the title compound, a Hirshfeld surface analysis was performed and the two-dimensional fingerprint plots were generated with Crystal Explorer17 (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). CrystalExplorer17. The University of Western Australia.]). The Hirshfeld surface mapped over dnorm and corresponding colours representing various inter­actions are shown in Fig. 4[link]. The red points on the surface correspond to the N—H⋯Cl and C—H⋯Cl inter­actions. The two-dimensional fingerprint plots (McKinnon et al., 2007[McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814-3816.]) are shown in Fig. 5[link]. On the Hirshfeld surface, the largest contribution (53.1%) comes from the weak van der Waals H⋯H contacts. The inter­action of dnorm on the two-dimensional fingerprint plot shows two spikes; each one corresponds to H⋯H (39%) and H⋯Cl/Cl⋯H (32.5%) respectively. The H⋯Cl inter­action highlights the hydrogen bond between adjacent moieties in the crystal structure. The C⋯H/H⋯C (16.5%) inter­actions appear as two shoulders. These inter­actions play a crucial role in the overall stabilization of the crystal packing.

[Figure 4]
Figure 4
Hirshfeld surface mapped over dnorm highlighting the regions of N—H⋯Cl and C—H⋯Cl inter­molecular contacts.
[Figure 5]
Figure 5
The full two-dimensional fingerprint plot for the organic cation in the title compound and those delineated into H⋯H (39%), Cl⋯H/H⋯Cl (32.5%) and C⋯H/H⋯C (16.5%) contacts.

4. Database survey

A search of the Cambridge Structural Database (CSD, Version 5.41, update of August 2020; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) revealed two related complexes containing 2-methyl­pyridinium: [2-methyl­pyridinium tetra­chloro­ferrate(III)] (CCDC refcode WAYJEN; Wyrzykowski et al., 2011[Wyrzykowski, D., Wera, M., Sikorski, A., Jacewicz, D. & Chmurzyński, L. (2011). Cent. Eur. J. Chem. 9, 1096-1101.]) and [bis­(2-methyl-pyridinium) tetra­chloro-zinc(II)] (CCDC refcode WIPCUW; Jasrotia et al., 2018[Jasrotia, D., Verma, S. K., Sridhar, B., Alvi, P. A. & Kumar, A. (2018). Mater. Chem. Phys. 207, 98-104.]). The mol­ecular structures of both WAYJEN and WIPCUW display three-dimensional supra­molecular networks arising from N—H⋯Cl and C—H⋯Cl inter­actions. In addition, the search also revealed a 2-methyl­pyridine and copper chloride complex: [di­chloro-bis­(2-methyl­pyridine)Cu(II)] (CCDC refcode CMPYCU01; Marsh et al., 1982[Marsh, W. E., Hatfield, W. E. & Hodgson, D. J. (1982). Inorg. Chem. 21, 2679-2684.]) and [aqua-di­chloro-bis­(2-methyl­pyridine)Cu(II)] (CCDC refcode BIJWUM; Marsh et al., 1982[Marsh, W. E., Hatfield, W. E. & Hodgson, D. J. (1982). Inorg. Chem. 21, 2679-2684.]) and a very recently published di­chlorido­methano­lbis(2-methyl­pyridine)Cu(II) complex (Reddy, 2020[Reddy, J. P. (2020). Acta Cryst. E76, 1771-1774.]). All of these structures display three-dimensional supra­molecular networks stabilized by C—H⋯Cl and O—H⋯Cl inter­actions.

5. Synthesis and crystallization

Both 2-methyl­pyridine and anhydrous copper(II) chloride obtained from Aldrich were used for the reaction. Anhydrous copper(II) chloride (0.495 g, 0.005 mol) was dissolved in 10 ml of distilled water. To this solution, 2-methyl­pyridine (0.93 g, 0.01 mol) was added followed by addition of few drops of HCl (36%) and the resulting mixture was stirred for ∼30 min. at room temperature. The solution was then allowed to stand at room temperature for a few hours before being filtered and left at room temperature for crystallization. Block-shaped, pale-yellow-coloured crystals were obtained after 36 h.

6. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. H atoms were placed in calculated positions with C—H = 0.93–0.96 Å and N—H = 0.88 Å and refined as riding with fixed isotropic displacement parameters [Uiso(H) = 1.2–1.5Ueq(C, N)].

Table 2
Experimental details

Crystal data
Chemical formula (C6H8N)2[CuCl4]
Mr 393.62
Crystal system, space group Monoclinic, I2/c
Temperature (K) 120
a, b, c (Å) 15.2354 (8), 8.3683 (3), 12.8372 (6)
β (°) 99.205 (5)
V3) 1615.60 (13)
Z 4
Radiation type Mo Kα
μ (mm−1) 2.00
Crystal size (mm) 0.32 × 0.27 × 0.25
 
Data collection
Diffractometer Agilent Xcalibur, Sapphire3
Absorption correction Analytical (CrysAlis PRO; Agilent, 2014[Agilent (2014). CrysAlis PRO. Agilent Technologies Ltd, Yarnton, England.])
Tmin, Tmax 0.848, 0.965
No. of measured, independent and observed [I > 2σ(I)] reflections 24556, 2821, 2324
Rint 0.071
(sin θ/λ)max−1) 0.758
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.036, 0.081, 1.08
No. of reflections 2821
No. of parameters 88
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.58, −0.58
Computer programs: CrysAlis PRO (Agilent, 2014[Agilent (2014). CrysAlis PRO. Agilent Technologies Ltd, Yarnton, England.]), SUPERFLIP (Palatinus & Chapuis, 2007[Palatinus, L. & Chapuis, G. (2007). J. Appl. Cryst. 40, 786-790.]; Palatinus & van der Lee, 2008[Palatinus, L. & van der Lee, A. (2008). J. Appl. Cryst. 41, 975-984.]; Palatinus et al., 2012[Palatinus, L., Prathapa, S. J. & van Smaalen, S. (2012). J. Appl. Cryst. 45, 575-580.]), 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: CrysAlis PRO (Agilent, 2014); cell refinement: CrysAlis PRO (Agilent, 2014); data reduction: CrysAlis PRO (Agilent, 2014); program(s) used to solve structure: Superflip (Palatinus & Chapuis, 2007; Palatinus & van der Lee, 2008; Palatinus et al., 2012); 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-methylpyridinium) tetrachloridocuprate(II) top
Crystal data top
(C6H8N)2[CuCl4]F(000) = 796
Mr = 393.62Dx = 1.618 Mg m3
Monoclinic, I2/cMo Kα radiation, λ = 0.71073 Å
a = 15.2354 (8) ÅCell parameters from 5462 reflections
b = 8.3683 (3) Åθ = 2.9–32.5°
c = 12.8372 (6) ŵ = 2.00 mm1
β = 99.205 (5)°T = 120 K
V = 1615.60 (13) Å3Blocks, pale yellow
Z = 40.32 × 0.27 × 0.25 mm
Data collection top
Agilent Xcalibur, Sapphire3
diffractometer
2324 reflections with I > 2σ(I)
Detector resolution: 16.1511 pixels mm-1Rint = 0.071
ω scansθmax = 32.6°, θmin = 2.9°
Absorption correction: analytical
(CrysAlisPro; Agilent, 2014)
h = 2222
Tmin = 0.848, Tmax = 0.965k = 1112
24556 measured reflectionsl = 1919
2821 independent reflections
Refinement top
Refinement on F2Primary atom site location: iterative
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.036H-atom parameters constrained
wR(F2) = 0.081 w = 1/[σ2(Fo2) + (0.0308P)2 + 1.4147P]
where P = (Fo2 + 2Fc2)/3
S = 1.08(Δ/σ)max < 0.001
2821 reflectionsΔρmax = 0.58 e Å3
88 parametersΔρmin = 0.58 e Å3
0 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
Cu10.5000000.22261 (4)0.2500000.01825 (9)
Cl10.38953 (3)0.10590 (5)0.13863 (3)0.02196 (11)
Cl20.41898 (3)0.32033 (5)0.36739 (4)0.02260 (11)
N10.68501 (11)0.01294 (19)0.63521 (12)0.0218 (3)
H10.6427340.0542960.6452100.026*
C20.66434 (12)0.1697 (2)0.62543 (13)0.0197 (3)
C30.73176 (13)0.2747 (2)0.61076 (14)0.0222 (4)
H30.7204580.3862730.6059790.027*
C40.83354 (14)0.0533 (2)0.61276 (16)0.0261 (4)
H40.8908160.0126220.6071220.031*
C50.81548 (14)0.2165 (2)0.60311 (15)0.0247 (4)
H50.8611020.2883800.5911780.030*
C60.76632 (14)0.0468 (2)0.63060 (15)0.0254 (4)
H60.7770880.1580920.6397040.031*
C70.57039 (13)0.2174 (2)0.62851 (16)0.0259 (4)
H7A0.5363950.2143430.5569600.039*
H7B0.5691410.3259760.6568450.039*
H7C0.5439360.1431890.6737670.039*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Cu10.02142 (16)0.01465 (15)0.01929 (16)0.0000.00510 (12)0.000
Cl10.0253 (2)0.0174 (2)0.0226 (2)0.00165 (16)0.00223 (17)0.00185 (15)
Cl20.0276 (2)0.0172 (2)0.0246 (2)0.00035 (16)0.00935 (17)0.00340 (16)
N10.0268 (8)0.0154 (7)0.0235 (7)0.0023 (6)0.0054 (6)0.0011 (6)
C20.0264 (9)0.0172 (8)0.0155 (8)0.0004 (7)0.0033 (7)0.0000 (6)
C30.0311 (10)0.0150 (8)0.0203 (8)0.0030 (7)0.0040 (7)0.0004 (6)
C40.0251 (10)0.0266 (10)0.0267 (9)0.0016 (8)0.0040 (8)0.0010 (8)
C50.0281 (10)0.0221 (9)0.0244 (9)0.0069 (7)0.0056 (8)0.0008 (7)
C60.0326 (10)0.0172 (9)0.0265 (10)0.0031 (7)0.0048 (8)0.0016 (7)
C70.0263 (10)0.0243 (10)0.0276 (10)0.0018 (7)0.0061 (8)0.0004 (8)
Geometric parameters (Å, º) top
Cu1—Cl1i2.2496 (5)C3—C51.383 (3)
Cu1—Cl12.2496 (5)C4—H40.9500
Cu1—Cl22.2492 (4)C4—C51.395 (3)
Cu1—Cl2i2.2492 (4)C4—C61.370 (3)
N1—H10.8800C5—H50.9500
N1—C21.350 (2)C6—H60.9500
N1—C61.346 (2)C7—H7A0.9800
C2—C31.387 (3)C7—H7B0.9800
C2—C71.493 (3)C7—H7C0.9800
C3—H30.9500
Cl1—Cu1—Cl1i128.54 (3)C5—C4—H4121.0
Cl2i—Cu1—Cl199.614 (17)C6—C4—H4121.0
Cl2i—Cu1—Cl1i98.550 (17)C6—C4—C5118.10 (19)
Cl2—Cu1—Cl198.549 (17)C3—C5—C4120.61 (18)
Cl2—Cu1—Cl1i99.616 (17)C3—C5—H5119.7
Cl2i—Cu1—Cl2137.36 (3)C4—C5—H5119.7
C2—N1—H1118.0N1—C6—C4119.91 (18)
C6—N1—H1118.0N1—C6—H6120.0
C6—N1—C2124.00 (17)C4—C6—H6120.0
N1—C2—C3117.46 (17)C2—C7—H7A109.5
N1—C2—C7117.88 (17)C2—C7—H7B109.5
C3—C2—C7124.64 (17)C2—C7—H7C109.5
C2—C3—H3120.1H7A—C7—H7B109.5
C5—C3—C2119.88 (17)H7A—C7—H7C109.5
C5—C3—H3120.1H7B—C7—H7C109.5
N1—C2—C3—C52.2 (3)C6—N1—C2—C30.6 (3)
C2—N1—C6—C41.5 (3)C6—N1—C2—C7177.92 (17)
C2—C3—C5—C41.7 (3)C6—C4—C5—C30.4 (3)
C5—C4—C6—N12.0 (3)C7—C2—C3—C5176.27 (18)
Symmetry code: (i) x+1, y, z+1/2.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N1—H1···Cl1ii0.882.933.4297 (16)118
N1—H1···Cl2ii0.882.413.2050 (16)150
C6—H6···Cl1iii0.952.623.453 (2)147
C7—H7A···Cl20.982.923.850 (2)159
C7—H7B···Cl2iv0.982.983.872 (2)151
Symmetry codes: (ii) x+1, y, z+1; (iii) x+1/2, y1/2, z+1/2; (iv) x+1, y+1, z+1.
 

Acknowledgements

We thank Dr J. S. Yadav, Provost and Director (R&D) for his support and encouragement.

Funding information

Funding for this research was provided by: Indrashil University.

References

First citationAdams, C. J., Crawford, P. C., Orpen, A. G., Podesta, T. J. & Salt, B. (2005). Chem. Commun. pp. 2457–2458.  Web of Science CSD CrossRef Google Scholar
First citationAgilent (2014). CrysAlis PRO. Agilent Technologies Ltd, Yarnton, England.  Google Scholar
First citationBurslem, G. M., Kyle, H. F., Prabhakaran, P., Breeze, A. L., Edwards, T. A., Warriner, S. L., Nelson, A. & Wilson, A. J. (2016). Org. Biomol. Chem. 14, 3782–3786.  CrossRef CAS PubMed Google Scholar
First citationDodds, C. A. & Kennedy, A. R. (2018). Acta Cryst. E74, 1369–1372.  Web of Science CSD CrossRef IUCr Journals 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 citationEddaoudi, M., Kim, J., Rosi, N., Vodak, D., Wachter, J., O'Keeffe, M. & Yaghi, O. M. (2002). Science, 295, 469–472.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationFérey, G., Latroche, M., Serre, C., Millange, F., Loiseau, T. & Percheron-Guégan, A. (2003). Chem. Commun. pp. 2976–2977.  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 citationJames, S. L. (2003). Chem. Soc. Rev. 32, 276–288.  Web of Science CrossRef PubMed CAS Google Scholar
First citationJasrotia, D., Verma, S. K., Sridhar, B., Alvi, P. A. & Kumar, A. (2018). Mater. Chem. Phys. 207, 98–104.  CSD CrossRef CAS Google Scholar
First citationKitagawa, S., Kitaura, R. & Noro, S. I. (2004). Angew. Chem. Int. Ed. 43, 2334–2375.  Web of Science CrossRef CAS Google Scholar
First citationMarsh, W. E., Hatfield, W. E. & Hodgson, D. J. (1982). Inorg. Chem. 21, 2679–2684.  CSD CrossRef CAS Web of Science Google Scholar
First citationMcKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814–3816.  Web of Science CrossRef Google Scholar
First citationMolano, M. F., Lorett Velasquez, V. P., Erben, M. F., Nossa González, D. L., Loaiza, A. E., Echeverría, G. A., Piro, O. E., Tobón, Y. A., Ben Tayeb, K. & Gómez Castaño, J. A. (2020). Acta Cryst. E76, 148–154.  CSD CrossRef IUCr Journals Google Scholar
First citationMoulton, B. & Zaworotko, M. J. (2001). Chem. Rev. 101, 1629–1658.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMueller, U., Schubert, M., Teich, F., Puetter, H., Schierle-Arndt, K. & Pastré, J. (2006). J. Mater. Chem. 16, 626–636.  Web of Science CrossRef CAS Google Scholar
First citationPalatinus, L. & Chapuis, G. (2007). J. Appl. Cryst. 40, 786–790.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPalatinus, L., Prathapa, S. J. & van Smaalen, S. (2012). J. Appl. Cryst. 45, 575–580.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPalatinus, L. & van der Lee, A. (2008). J. Appl. Cryst. 41, 975–984.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPrakashaReddy, J. & Pedireddi, V. R. (2007). Eur. J. Inorg. Chem. pp. 1150–1158.  Web of Science CSD CrossRef Google Scholar
First citationReddy, J. P. (2020). Acta Cryst. E76, 1771–1774.  CSD CrossRef IUCr Journals Google Scholar
First citationReddy, J. P., Swain, D. & Pedireddi, V. R. (2014). Cryst. Growth Des. 14, 5064–5071.  CSD CrossRef CAS Google Scholar
First citationRuben, M., Ziener, U., Lehn, J. M., Ksenofontov, V., Gütlich, P. & Vaughan, G. B. M. (2005). Chem. Eur. J. 11, 94–100.  Web of Science CSD CrossRef Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationStavila, V., Talin, A. A. & Allendorf, M. D. (2014). Chem. Soc. Rev. 43, 5994–6010.  Web of Science CrossRef CAS PubMed Google Scholar
First citationThorat, V. H., Ingole, T. S., Vijayadas, K. N., Nair, R. V., Kale, S. S., Ramesh, V. V. E., Davis, H. C., Prabhakaran, P., Gonnade, R. G., Gawade, R. L., Puranik, V. G., Rajamohanan, P. R. & Sanjayan, G. J. (2013). Eur. J. Org. Chem. pp. 3529–3542.  CSD CrossRef Google Scholar
First citationTurner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). CrystalExplorer17. The University of Western Australia.  Google Scholar
First citationVittaya, L., Leesakul, N., Saithong, S. & Chainok, K. (2015). Acta Cryst. E71, m201–m202.  CSD CrossRef IUCr Journals Google Scholar
First citationWan, Y., Yang, H. & Zhao, D. (2006). Acc. Chem. Res. 39, 423–432.  Web of Science CrossRef PubMed CAS Google Scholar
First citationWyrzykowski, D., Wera, M., Sikorski, A., Jacewicz, D. & Chmurzyński, L. (2011). Cent. Eur. J. Chem. 9, 1096–1101.  CAS Google Scholar
First citationZiach, K., Chollet, C., Parissi, V., Prabhakaran, P., Marchivie, M., Corvaglia, V., Bose, P. P., Laxmi-Reddy, K., Godde, F., Schmitter, J.-M., Chaignepain, S., Pourquier, P. & Huc, I. (2018). Nat. Chem. 10, 511–518.  Web of Science CSD CrossRef CAS PubMed 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