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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structure, spectroscopic characterization and Hirshfeld surface analysis of aqua­di­chlorido­{N-[(pyridin-2-yl)methyl­­idene]aniline}copper(II) monohydrate

CROSSMARK_Color_square_no_text.svg

aLaboratorio de Química Teórica y Computacional, Grupo Química-Física Molecular y Modelamiento Computacional (QUIMOL), Facultad de Ciencias, Universidad Pedagógica y Tecnológica de Colombia, Tunja, Boyacá, 050030, Colombia, bUniversidad Antonio Nariño, Facultad de Ciencias, Bogotá, Colombia, cCentro de Química Inorgánica (CEQUINOR), Facultad de Ciencias Exactas, Universidad Nacional de La Plata, C.C. 962, 1900 La Plata, Argentina, dDepartamento de Química, Facultad de Ciencias, Pontificia Universidad Javeriana, 110231561 Bogotá, Colombia, eDepartamento de Física, Facultad de Ciencias Exactas, Universidad Nacional de La Plata and IFLP(CONICET), C.C. 67, 1900 La Plata, Argentina, and fUniversity of Lille, CNRS, UMR 8516, LASIR - Laboratoire de Spectrochimie Infrarouge et Raman, F-59000 Lille, France
*Correspondence e-mail: jovanny.gomez@uptc.edu.co

Edited by J. Ellena, Universidade de Sâo Paulo, Brazil (Received 21 November 2019; accepted 24 December 2019; online 7 January 2020)

The reaction of N-phenyl-1-(pyridin-2-yl)methanimine with copper chloride dihydrate produced the title neutral complex, [CuCl2(C12H10N2)(H2O)]·H2O. The CuII ion is five-coordinated in a distorted square-pyramidal geometry, in which the two N atoms of the bidentate Schiff base, as well as one chloro and a water mol­ecule, form the irregular base of the pyramidal structure. Meanwhile, the apical chloride ligand inter­acts through a strong hydrogen bond with a water mol­ecule of crystallization. In the crystal, mol­ecules are arranged in pairs, forming a stacking of symmetrical cyclic dimers that inter­act in turn through strong hydrogen bonds between the chloride ligands and both the coordinated and the crystallization water mol­ecules. The mol­ecular and electronic structures of the complex were also studied in detail using EPR (continuous and pulsed), FT–IR and Raman spectroscopy, as well as magnetization measurements. Likewise, Hirshfeld surface analysis was used to investigate the inter­molecular inter­actions in the crystal packing.

1. Chemical context

CuII ions coordinated by di­imine N-donor ligands (–N=C—C=N–) are of great inter­est since they combine structural flexibility with other desired characteristics, such as ease of preparation, photophysical (Barwiolek et al., 2016[Barwiolek, M., Szczęsny, R. & Szłyk, E. (2016). J. Chem. Sci. 128, 1057-1066.]) and photobiological (Banerjee et al., 2016[Banerjee, S., Dixit, A., Maheswaramma, K. S., Maity, B., Mukherjee, S., Kumar, A., Karande, A. A. & Chakravarty, A. R. (2016). J. Chem. Sci. 128, 165-175.]) properties, and catalytic activity (Dias et al., 2010[Dias, P. M., Kinouti, L., Constantino, V. R. L., Ferreira, A. M. D. C., Gonçalves, M. B., Nascimento, R. R., Petrilli, H. M., Caldas, M. & Frem, R. C. G. (2010). Quím. Nova, 33, 2135-2142.]), as well as the capability to mimic active protein sites (Gupta & Sutar, 2008[Gupta, K. C. & Sutar, A. K. (2008). Coord. Chem. Rev. 252, 1420-1450.]) and stabilize both metal oxidation states common in biological systems. These complexes also exhibit a broad spectrum of pharmacological properties including anti-inflammatory, anti­bacterial, anti­oxidant and anti­metastatic (Chaviara et al., 2005[Chaviara, A. Th., Christidis, P. C., Papageorgiou, A., Chrysogelou, E., Hadjipavlou-Litina, D. J. & Bolos, C. A. (2005). J. Inorg. Biochem. 99, 2102-2109.]) activities. In particular, they are promising metallotherapeutic drugs for the treatment of cancer, given their ability to induce apoptosis or generate reactive oxygen species (ROS) in oxidative stress, resulting in DNA damage and strand breaks in cancerous cells (Trudu et al., 2015[Trudu, F., Amato, F., Vaňhara, P., Pivetta, T., Peña-Méndez, E. M. & Havel, J. (2015). J. Appl. Biomed. 13, 79-103.]).

In particular, bidentate pyridinyl­imine (C5H4N—CH2—NH—C6H5) and pyridinyl­methyl­amine (C5H4N–CH=N—C6H5) Schiff base derivatives have attracted increasing attention because of their close structural relationship with the protein Aβ aggregate p-I-stilbene [I-C6H4—CH=N—C6H4R, R = N(CH3)2] and thus their potential use for the development of metal chelators for the attenuation of metal-involved neurodegeneration in Alzheimer's disease (DeToma et al., 2012[DeToma, A. S., Salamekh, S., Ramamoorthy, A. & Lim, M. H. (2012). Chem. Soc. Rev. 41, 608-621.]). These ligands can therefore act as chemical reagents that can target metal-associated amyloid-β (Aβ) species and modulate metal-induced Aβ aggregation and neurotoxicity in vitro and in living cells (Braymer et al., 2012[Braymer, J. J., Merrill, N. M. & Lim, M. H. (2012). Inorg. Chim. Acta, 380, 261-268.]).

Based on their relevant structural features and promising biological activity, we have begun to explore novel metal complexes coordinated with di­imine ligands (Schiff bases). We report here the synthesis and structural characterization of the complex [Cu(H2O)Cl2(C12H10N2)]·H2O where C12H10N2 = N-(pyridin-2-yl­methyl­ene)aniline. This compound is formed by the reaction of copper chloride dihydrate with the C12H10N2 Schiff base to afford bright-green crystals suitable for X-ray diffraction studies.

[Scheme 1]

2. Structural commentary

The title complex crystallizes in the monoclinic space group P21/n with Z = 4 mol­ecules per unit cell. The CuII ion is five-coordinated by two nitro­gen atoms from the di­imine ligand and a water mol­ecule in the equatorial position, and two chloro ligands that provide an apical and a pseudo-equatorial coordination, as shown in Fig. 1[link]. The apical Cu1—Cl2 distance is 0.193 Å longer than that of the non-apical Cu1—Cl1 distance (Table 1[link]). The apical chloro atom, Cl2, is hydrogen bonded to the water mol­ecule of crystallization, O2W—H2A⋯Cl2 [2.357 (12) Å; Table 2[link]]. As a result, the coordination geometry around the CuII ion is best described as a distorted square-pyramidal structure with a trigonal–bipyramidal component of structural index τ = 0.40 [= (β - α)/60, where β = O1W—Cu1—N2 = 169.12 (8)° and α = N1—Cu1—Cl1 = 145.15 (6)°]; for perfect square-pyramidal and trigonal–bipyramidal geom­etries, the values of τ are zero and unity, respectively (Addison et al., 1984[Addison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349-1356.]). The two N—Cu distances, however, differ by only 0.019 Å. The trigonal-component axial compression (%TC) is −3.22 [%TC = 100(BD)/B, where B is the Cu1—O1W distance and D the Cu1—N1 distance; Addison et al., 1984[Addison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349-1356.]]. Intra­molecular C8—H6⋯Cl2 and C5—H4⋯O1W hydrogen bonds occur (Table 2[link]).

Table 1
Selected geometric parameters (Å, °)

N1—Cu1 2.046 (2) Cl1—Cu1 2.2744 (7)
N2—Cu1 2.0345 (19) Cl2—Cu1 2.4673 (7)
Cu1—O1W 1.9821 (18) C7—N2 1.431 (3)
       
O1W—Cu1—N2 169.12 (8) N1—Cu1—Cl1 145.15 (6)
O1W—Cu1—N1 88.66 (8) O1W—Cu1—Cl2 88.86 (6)
N2—Cu1—N1 80.47 (8) N2—Cu1—Cl2 94.66 (6)
O1W—Cu1—Cl1 92.92 (6) N1—Cu1—Cl2 106.80 (6)
N2—Cu1—Cl1 95.76 (6) Cl1—Cu1—Cl2 108.04 (3)

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O1W—H1A⋯O2Wi 0.85 (1) 1.80 (1) 2.640 (3) 171 (3)
O1W—H1B⋯Cl2ii 0.86 (1) 2.35 (2) 3.1740 (19) 162 (3)
O2W—H2A⋯Cl2 0.85 (1) 2.36 (1) 3.199 (2) 171 (3)
O2W—H2B⋯Cl1ii 0.85 (1) 2.38 (1) 3.215 (2) 167 (3)
C5—H4⋯O1W 0.95 (2) 2.47 (2) 2.999 (3) 115.3 (18)
C6—H5⋯Cl1iii 1.00 (3) 2.56 (3) 3.540 (3) 169 (2)
C8—H6⋯Cl2 0.87 (3) 2.95 (3) 3.707 (3) 147 (2)
C12—H10⋯Cl2iv 0.95 (3) 2.85 (3) 3.667 (3) 145 (2)
Symmetry codes: (i) x, y+1, z; (ii) -x+2, -y+1, -z+1; (iii) [-x+{\script{3\over 2}}, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iv) [-x+{\script{3\over 2}}, y+{\script{1\over 2}}, -z+{\script{1\over 2}}].
[Figure 1]
Figure 1
ORTEP representation of the title complex with the atom numbering. Displacement ellipsoids are drawn at 50% probability level. The hydrogen bond to the water mol­ecule of crystallization is shown as a dashed blue line.

3. Supra­molecular features

As expected, both the water mol­ecule of crystallization and the aqua ligand play a significant role in the crystal packing of the complex. This is governed by the presence of symmetric cyclic dimers formed between complex mol­ecules in adjacent unit cells along the a-axis direction (see Fig. 2[link]). Each dimer comprises two water mol­ecules of crystallization, each of which links the two complex monomers by two different hydrogen bonds, one with the apical Cl2 ligand [2.357 (12) Å] and the second with the non-apical Cl1 ligand [2.384 (13) Å]. The dimers are stacked alternately in the b-axis direction, forming a wave-like arrangement as shown in Fig. 3[link]. Each dimer inter­acts with two other dimers through two different hydrogen bonds. One of these [2.85 (3) Å], is formed between the apical chlorine Cl2 and the aromatic hydrogen H10 in the ortho position, while the second [2.56 (3) Å], is formed between the non-apical chlorine Cl1 and the aliphatic hydrogen H5 of the CH group. In addition, each apical chloro ligand Cl2 inter­acts with the hydrogen atom H1B of the aqua O atom O1W of a third dimer by means of a shorter hydrogen bond [2.347 (15) Å], with the other hydrogen atom H2A of the aqua ligand, forming a quite short hydrogen bond [1.795 (11) Å] with the oxygen atom O2W of the water mol­ecule of crystallization.

[Figure 2]
Figure 2
Symmetric cyclic dimers in the crystal structure of the title complex formed by dual Cl⋯H—O—H⋯Cl inter­actions (dotted blue lines) between the chlorine ligands and the water mol­ecules.
[Figure 3]
Figure 3
The crystal packing in a view along the a + b vector showing the stacking of symmetric cyclic dimers.

4. Hirshfeld surface analysis

In order to investigate and visualize the role of weak inter­molecular inter­actions, a Hirshfeld surface (HS) analysis (Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]) was carried out and the associated two-dimensional fingerprint plots (McKinnon et al., 2007[McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814-3816.]) generated using CrystalExplorer17.5 (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. University of Western Australia.]). The three dimensional dnorm surface of the title compound using a standard surface resolution with a fixed colour scale of 0.5 to 1.5 a.u. is shown in Fig. 4[link]. The darkest red spots on this surface correspond to the H2O⋯H—O and H—O—H⋯Cl hydrogen bonds resulting from the inter­actions between the water mol­ecule of crystallization and the coordinated water and chlorine, respectively. The fingerprint plots in Fig. 5[link], for all inter­actions in the title compound, and those delineated into H⋯H, Cl⋯H, C⋯H, H⋯O/O⋯H and N⋯H contacts, exhibit four pseudo-symmetric long sharp spikes characteristic of strong hydrogen bonds and one spike in the de and di diagonal axes associated with H⋯H inter­actions. The greatest contribution to the HS is from H⋯H inter­actions (44.0%), which are represented by a distinctive sharp spike in the region de = di ≃ 1.5 Å. The Cl⋯H contacts make a 23.1% contribution to the HS and are represented by a pair of sharp spikes in the region de + di ≃ 2.2 Å. The C⋯H contacts (13.2% contribution) are observed as two wide contour signals in the region de + di ≃ 3.0 Å. The N⋯H contacts (2.7%) are represented by two signals with thick edges in the region de + di ≃ 3.3 Å. The O⋯H contacts are represented by two sharp spikes in the region de + di ≃ 1.6 Å, which indicates a clear formation of hydrogen bonds.

[Figure 4]
Figure 4
Hirshfeld surface of the title complex mapped over electrostatic potential in the range 0.5 to 1.5 atomic units.
[Figure 5]
Figure 5
Two-dimensional fingerprint plots for all contacts and delineated into H⋯H, Cl⋯H, C⋯H, H⋯O/O⋯H, and N⋯H contacts in the title complex.

5. CW-EPR/Pulsed-EPR and PPMS characterization

In order to obtain in-depth information on the spin properties of this unpaired spin complex (d9, 2S + 1 = 2), electron paramagnetic resonance (EPR) continuous-wave (CW) experiments were performed on a X-Band Bruker ELEXSYS E500 spectrometer operating at 9.8 GHz. The powdered sample was inserted in a quartz tube and the spectra were recorded at room temperature and 100 K under non-saturated conditions: microwave power of 0.63 mW and modulation amplitude of 2 G. Pulsed EPR was studied at 5 K with a Bruker ELEXSYS E580 spectrometer equipped with a helium flow cryostat. Two-pulse echo field sweep acquisitions were performed using a standard Hahn echo sequence 90 − τ − 180 with a 90° pulse length of 16 ns and τ value of 172 ns. The HYperfine Sublevel CORrElation spectroscopy (HYSCORE) experiments (Höfer et al., 1986[Höfer, P., Grupp, A., Nebenführ, H. & Mehring, M. (1986). Chem. Phys. Lett. 132, 279-282.]) were recorded with 256 × 256 data points for both the t1 and t2 time domains, 90° pulse length of 16 ns and an echo delay of 172 and 200 ns. The obtained HYSCORE spectra are composed of two quadrants: the first quadrant (+,−) where A > 2nI (nI being the nuclear frequency) corresponds to the strong hyperfine coupling A between the I nucleus and the unpaired electron and the second quadrant (+,+) where A < 2nI corresponds to weaker inter­actions. Magnetization measurements were performed with a physical property measurement system (PPMS) Quantum Design Dynacool of 9 T and the vibrating sample magnetometer (VSM) option. To verify that both samples, i.e. powder and crystal, correspond to the same compound, a comparison between the X-ray powder diffraction pattern and the simulated single X-ray diffraction pattern is presented in Fig. 6[link].

[Figure 6]
Figure 6
Comparison between the experimental and theoretical powder diffractogram for the title complex. The calculated diffractogram was simulated using Mercury software (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. In the press. https://doi.org/10.1107/S1600576719014092]) from the CIF file.

A strong isotropic signal characteristic of CuII was detected by EPR spectroscopy with giso = 2.13 and a line width of 130 G at room temperature. No sign of anisotropic behaviour was detected in the low-temperature continuous wave EPR spectra recorded at 100 K (Fig. 7[link]) and 8 K. Further efforts to reveal possible minor anisotropic behavior through Q-band (34 GHz) low-temperature (8 K) measurements still showed a single strong band characteristic of the CuII ion in an isotropic environment. The isotropic signal was so dominant in the powder spectrum that the anisotropic features were invisible. A similar observation was made by Xavier & Murugesan (1998[Xavier, V. S. & Murugesan, A. R. (1998). Mol. Phys. 94, 269-273.]). 98 mg of CuII were qu­anti­fied in the complex sample by continuous wave EPR using copper sulfate with a known mass as standard. To obtain more information about the surroundings of the copper(II) centre, a pulsed EPR experiment was performed using HYSCORE. In quadrant (+,+), a unique signal characteristic of hydrogen was detected with a Larmor frequency of 14.6 MHz corresponding to a weak inter­action between the copper unpaired electron and the hydrogen nucleus (Fig. 8[link]). Inter­actions between the copper and the nitro­gen atoms were not observed. It is probable that the inter­action is too strong to be detected by the HYSCORE sequence.

[Figure 7]
Figure 7
CW–EPR spectrum for the title complex measured at 100 K.
[Figure 8]
Figure 8
Two-dimensional HYSCORE spectrum of the title complex recorded at 5 K.

Magnetic susceptibility measurements were performed to verify the nature of coupling between the cupric ions. The temperature dependence of the molar magnetic susceptibility χM and the corresponding inverse susceptibility 1/χM measured at a magnetic field of 0.1 T in the temperature range of 2–400 K is shown in Fig. 9[link]. Fig. 10[link] shows the dependence of magnetization on the magnetic field at 2 K, 100 K and 300 K. At higher temperature, the magnetization manifest Curie–Weiss-like behaviour. The magnetization curves of the sample have features typical of a paramagnetic contribution between magnetic centers

[Figure 9]
Figure 9
Temperature dependence of molar magnetic susceptibility χM and inverse susceptibility 1/χM.
[Figure 10]
Figure 10
Dependence of magnetization on the magnetic field at 2, 100 and 300 K.

6. Database survey

A survey of the Cambridge Structural Database (CSD, Version 5.40, Oct 2019; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) reveals that crystal structures have been reported for coordinated CuII and ZnII complexes containing N-(pyridin-2-ylmeth­yl)aniline and its deprotonated form N-(pyridin-2-yl­methyl­ene)aniline, respectively (Braymer et al., 2012[Braymer, J. J., Merrill, N. M. & Lim, M. H. (2012). Inorg. Chim. Acta, 380, 261-268.]). For the former complex, [Cu(C12H12N2)Cl2], a nearly square-planar geometry between the bidentate Schiff base and two chloro ligands was reported; while for the latter, [Zn(C12H10N2)Cl2], a distorted tetra­hedral geometry was observed. A detailed revision of the CIF file reported for the [Cu(C12H12N2)Cl2] complex, however, reveals that the crystal packing of this compound can be best described as comprising polymeric chains of complex units consisting of slightly distorted square-pyramidal [(μ)Cu-Cu(C12H12N2)Cl2] where the apical position is occupied by a bridged CuII ion.

7. Synthesis and crystallization

The N-(pyridin-2-yl­methyl­ene)aniline ligand, C12H10N2, was prepared by condensation reaction between 2-pyridine­aldehyde (Sigma–Aldrich, 99%) and aniline (Sigma–Aldrich, 99%) in dry methanol (Merck, HPLC grade) at reflux temperature for 4 h under atmospheric pressure and constant stirring. The stoichiometry used in this reaction was 1:1 mmol. The released water vapour was prevented from returning to the reaction vessel by placing a condensation trap containing methanol in the lower base of the reflux column. No byproduct was formed during the reaction. The purity and mol­ecular weight of the ligand was confirmed by GC/MS spectrometry using an Agilent 6850 series II gas chromatograph (CG) coupled to an Agilent 5975B VL MSD mass spectrometer (MS) equipped with a split/splitess injection port (533 K, split ratio 15:1), with an Agilent 6850 series automatic injector and an Agilent 19091S-433E HP-5MS column.

MS (m/z ratio, %): C12H10N2 (182.1, 33.09%); C12H9N2 (181.1, 100%); C5H6 (77.05, 35.85%); C5H4N (78.05, 12.98%); C6H5N2 (105.05, 9.11%); C6H5N (91.05, 4.25%).

The title CuII complex was prepared by reacting the C12H10N2 ligand with copper chloride dihydrate, CuCl2·2H2O (Merck, 99.9%), in dry methanol for 8 h at room temperature, under atmospheric pressure and constant stirring. The CuII complex precipitated in methanol as a green solid, which was then separated from the solvent by rotoevaporation. Crystallization was carried out from a saturated solution of the CuII complex in methanol at 313 K, which was allowed to cool to room temperature and then hexane was added until reaching a 1:1 methanol/hexane ratio, followed by storage at 277 K. Crystals of the title CuII complex were separated by subtled deca­ntation and evaporation of the solvents at room temperature.

An infrared (IR) spectrum in attenuated total reflectance (ATR) was acquired from a ground crystal using a Shimadzu Prestigie-21 spectrophotometer with Fourier Transform (FTIR), equipped with a Michelson-type inter­ferometer, a KBr/Ge beam-splitter, a ceramic lamp and DLATGS detector. The FTIR spectrum was measured in the 4000–500 cm−1 range with a resolution of 3.0 cm−1 and 30 scans. Likewise, Raman spectra of the title complex were obtained using a LabRAM HR confocal Raman microscope (Horiba Scientific) operating in a backlit orientation and equipped with a cryogenic detector and laser lines of 473, 532 and 633 nm of 18, 30 and 17 mW maximum power, respectively. The micro-Raman spectra of the complex were taken through an Olympus 50× long-working-distance microscope objective (NA = 0.5, WD = 10.6 mm), in the range from 3500 to 100 cm−1, with a resolution of 4 cm−1 and a laser power of around 3.0 mW.

IR (ATR, cm−1): 3067 C—H Pyr, 3213 C—H Ph, 1601 C=N Schiff B., 1540 C=N Pyr, 1434 C—H, 693 Cu—N. Raman (cm−1), 1601 C=N Schiff B., 1540 C=N Pyr, 552 Cu—N, 693 Cu—N, 411 Cu—O, 272 Cu—Cl.

8. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. The hydrogen atoms were found in difference-Fourier maps. The O⋯H and H⋯H distances in both water molecules were fixed at 0.86 (1) and 1.36 (2) Å, respectively.

Table 3
Experimental details

Crystal data
Chemical formula [CuCl2(C12H10N2)(H2O)]·H2O
Mr 352.69
Crystal system, space group Monoclinic, P21/n
Temperature (K) 297
a, b, c (Å) 9.3322 (2), 7.7341 (2), 20.1143 (4)
β (°) 90.002 (2)
V3) 1451.77 (6)
Z 4
Radiation type Mo Kα
μ (mm−1) 1.87
Crystal size (mm) 0.24 × 0.12 × 0.07
 
Data collection
Diffractometer Rigaku Xcalibur Eos Gemini
Absorption correction Multi-scan (CrysAlis PRO; Rigaku OD, 2015[Rigaku OD (2015). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.])
Tmin, Tmax 0.888, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 6613, 3143, 2380
Rint 0.027
(sin θ/λ)max−1) 0.678
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.035, 0.080, 1.04
No. of reflections 3143
No. of parameters 228
No. of restraints 6
H-atom treatment All H-atom parameters refined
Δρmax, Δρmin (e Å−3) 0.26, −0.39
Computer programs: CrysAlis PRO (Rigaku OD, 2015[Rigaku OD (2015). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.]), SHELXT (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2014 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), ORTEP-3 for Windows (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]) and 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. In the press. https://doi.org/10.1107/S1600576719014092]).

Supporting information


Computing details top

Data collection: CrysAlis PRO (Rigaku OD, 2015); cell refinement: CrysAlis PRO (Rigaku OD, 2015); data reduction: CrysAlis PRO (Rigaku OD, 2015); program(s) used to solve structure: SHELXT (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2014 (Sheldrick, 2015b); molecular graphics: ORTEP-3 for Windows (Farrugia, 2012), Mercury (Macrae et al., 2020); software used to prepare material for publication: SHELXL2014 (Sheldrick, 2015b).

Aquadichlorido{N-[(pyridin-2-yl)methylidene]aniline}copper(II) monohydrate top
Crystal data top
[CuCl2(C12H10N2)(H2O)]·H2OF(000) = 716
Mr = 352.69Dx = 1.614 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 9.3322 (2) ÅCell parameters from 2118 reflections
b = 7.7341 (2) Åθ = 4.0–27.5°
c = 20.1143 (4) ŵ = 1.87 mm1
β = 90.002 (2)°T = 297 K
V = 1451.77 (6) Å3Fragment, green
Z = 40.24 × 0.12 × 0.07 mm
Data collection top
Rigaku Xcalibur Eos Gemini
diffractometer
3143 independent reflections
Radiation source: fine-focus sealed X-ray tube, Enhance (Mo) X-ray Source2380 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.027
Detector resolution: 16.0604 pixels mm-1θmax = 28.8°, θmin = 3.3°
ω scansh = 1210
Absorption correction: multi-scan
(CrysAlis PRO; Rigaku OD, 2015)
k = 109
Tmin = 0.888, Tmax = 1.000l = 2626
6613 measured reflections
Refinement top
Refinement on F2Primary atom site location: dual
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.035Hydrogen site location: difference Fourier map
wR(F2) = 0.080All H-atom parameters refined
S = 1.04 w = 1/[σ2(Fo2) + (0.031P)2 + 0.1405P]
where P = (Fo2 + 2Fc2)/3
3143 reflections(Δ/σ)max = 0.001
228 parametersΔρmax = 0.26 e Å3
6 restraintsΔρmin = 0.39 e Å3
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
C10.5545 (3)0.5653 (4)0.33174 (13)0.0375 (6)
C20.4122 (3)0.5200 (5)0.32604 (16)0.0548 (9)
C30.3235 (3)0.5432 (5)0.37988 (16)0.0551 (9)
C40.3789 (3)0.6103 (4)0.43668 (16)0.0482 (8)
C50.5233 (3)0.6512 (4)0.43961 (14)0.0409 (7)
C60.6568 (3)0.5473 (4)0.27790 (13)0.0399 (7)
C70.8855 (3)0.5724 (3)0.23313 (12)0.0340 (6)
C81.0201 (3)0.5067 (5)0.24304 (15)0.0538 (9)
C91.1138 (4)0.4932 (6)0.18965 (17)0.0664 (11)
C101.0737 (4)0.5471 (5)0.12820 (17)0.0616 (10)
C110.9387 (4)0.6116 (5)0.11776 (15)0.0569 (9)
C120.8442 (3)0.6267 (4)0.17029 (13)0.0432 (7)
N10.6108 (2)0.6281 (3)0.38848 (10)0.0326 (5)
N20.7884 (2)0.5841 (3)0.28799 (10)0.0305 (5)
Cl11.01089 (7)0.84014 (9)0.35841 (3)0.03986 (18)
Cl20.93385 (8)0.37596 (9)0.41578 (4)0.0452 (2)
Cu10.82853 (3)0.65623 (4)0.38345 (2)0.03006 (11)
O1W0.8277 (2)0.7241 (3)0.47848 (9)0.0375 (4)
H1A0.833 (3)0.8331 (13)0.4836 (13)0.040 (9)*
H1B0.896 (3)0.679 (3)0.5008 (16)0.089 (13)*
O2W0.8297 (2)0.0581 (3)0.50539 (11)0.0439 (5)
H2A0.854 (3)0.136 (3)0.4777 (11)0.072 (12)*
H2B0.883 (3)0.074 (4)0.5390 (10)0.082 (12)*
H10.383 (3)0.474 (4)0.2868 (14)0.047 (8)*
H20.231 (3)0.523 (4)0.3756 (15)0.064 (10)*
H30.326 (3)0.634 (4)0.4741 (14)0.049 (9)*
H40.564 (3)0.698 (3)0.4789 (12)0.033 (7)*
H50.615 (3)0.502 (4)0.2358 (14)0.051 (8)*
H61.041 (3)0.471 (4)0.2830 (15)0.056 (9)*
H71.199 (4)0.443 (4)0.1993 (16)0.074 (11)*
H81.135 (4)0.538 (5)0.0938 (18)0.082 (12)*
H90.908 (4)0.652 (4)0.0729 (17)0.079 (11)*
H100.751 (3)0.675 (3)0.1664 (13)0.042 (8)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
C10.0279 (13)0.0484 (17)0.0362 (14)0.0014 (12)0.0025 (11)0.0019 (13)
C20.0303 (16)0.084 (3)0.0495 (18)0.0072 (16)0.0085 (14)0.0086 (19)
C30.0248 (15)0.075 (3)0.066 (2)0.0036 (16)0.0003 (15)0.0034 (18)
C40.0336 (16)0.059 (2)0.0521 (19)0.0092 (14)0.0097 (15)0.0070 (17)
C50.0377 (16)0.0492 (18)0.0359 (15)0.0030 (13)0.0033 (12)0.0021 (14)
C60.0345 (15)0.0519 (18)0.0334 (14)0.0035 (13)0.0063 (12)0.0098 (14)
C70.0317 (14)0.0413 (15)0.0290 (13)0.0008 (12)0.0013 (11)0.0093 (12)
C80.0375 (17)0.089 (3)0.0347 (16)0.0156 (17)0.0055 (13)0.0090 (18)
C90.0364 (18)0.105 (3)0.057 (2)0.021 (2)0.0034 (16)0.018 (2)
C100.050 (2)0.088 (3)0.0474 (19)0.0040 (19)0.0170 (16)0.022 (2)
C110.063 (2)0.073 (2)0.0345 (16)0.0011 (18)0.0084 (15)0.0008 (17)
C120.0407 (17)0.0539 (19)0.0350 (15)0.0063 (15)0.0014 (13)0.0003 (14)
N10.0291 (11)0.0379 (13)0.0308 (11)0.0014 (10)0.0016 (9)0.0009 (10)
N20.0255 (11)0.0380 (12)0.0280 (11)0.0019 (9)0.0026 (9)0.0033 (10)
Cl10.0365 (4)0.0499 (4)0.0331 (3)0.0119 (3)0.0039 (3)0.0058 (3)
Cl20.0549 (4)0.0354 (4)0.0453 (4)0.0004 (3)0.0182 (3)0.0042 (3)
Cu10.02761 (18)0.0388 (2)0.02381 (17)0.00335 (14)0.00251 (12)0.00170 (13)
O1W0.0409 (11)0.0437 (13)0.0280 (9)0.0050 (10)0.0040 (8)0.0017 (9)
O2W0.0431 (12)0.0488 (13)0.0399 (11)0.0023 (10)0.0004 (10)0.0011 (11)
Geometric parameters (Å, º) top
C1—N11.347 (3)C8—C91.389 (4)
C1—C21.378 (4)C9—C101.357 (5)
C1—C61.450 (4)C10—C111.371 (5)
C2—C31.375 (4)C11—C121.381 (4)
C3—C41.357 (4)N1—Cu12.046 (2)
C4—C51.386 (4)N2—Cu12.0345 (19)
C5—N11.325 (3)Cu1—O1W1.9821 (18)
C6—N21.276 (3)Cl1—Cu12.2744 (7)
C7—C121.387 (4)Cl2—Cu12.4673 (7)
C7—C81.369 (4)C7—N21.431 (3)
N1—C1—C2122.6 (3)C1—N1—C5117.8 (2)
N1—C1—C6114.2 (2)C1—N1—Cu1112.58 (17)
C2—C1—C6123.2 (2)C5—N1—Cu1129.49 (17)
C1—C2—C3118.8 (3)C6—N2—C7118.2 (2)
C4—C3—C2118.9 (3)C6—N2—Cu1112.85 (18)
C3—C4—C5119.6 (3)C7—N2—Cu1128.92 (15)
N1—C5—C4122.4 (3)O1W—Cu1—N2169.12 (8)
N2—C6—C1119.5 (2)O1W—Cu1—N188.66 (8)
C12—C7—C8120.0 (3)N2—Cu1—N180.47 (8)
C12—C7—N2120.5 (2)O1W—Cu1—Cl192.92 (6)
C8—C7—N2119.5 (2)N2—Cu1—Cl195.76 (6)
C7—C8—C9119.5 (3)N1—Cu1—Cl1145.15 (6)
C10—C9—C8120.5 (3)O1W—Cu1—Cl288.86 (6)
C9—C10—C11120.3 (3)N2—Cu1—Cl294.66 (6)
C10—C11—C12120.0 (3)N1—Cu1—Cl2106.80 (6)
C7—C12—C11119.6 (3)Cl1—Cu1—Cl2108.04 (3)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O1W—H1A···O2Wi0.85 (1)1.80 (1)2.640 (3)171 (3)
O1W—H1B···Cl2ii0.86 (1)2.35 (2)3.1740 (19)162 (3)
O2W—H2A···Cl20.85 (1)2.36 (1)3.199 (2)171 (3)
O2W—H2B···Cl1ii0.85 (1)2.38 (1)3.215 (2)167 (3)
C5—H4···O1W0.95 (2)2.47 (2)2.999 (3)115.3 (18)
C6—H5···Cl1iii1.00 (3)2.56 (3)3.540 (3)169 (2)
C8—H6···Cl20.87 (3)2.95 (3)3.707 (3)147 (2)
C12—H10···Cl2iv0.95 (3)2.85 (3)3.667 (3)145 (2)
Symmetry codes: (i) x, y+1, z; (ii) x+2, y+1, z+1; (iii) x+3/2, y1/2, z+1/2; (iv) x+3/2, y+1/2, z+1/2.
 

Acknowledgements

The authors greatly acknowledge the financial support provided by the host institutions, i.e., Universidad Pedagógica y Tecnológica de Colombia (UPTC), Universidad António Nariño, Pontificia Universidad Javeriana, Universidad Nacional de La Plata (UNLP), Argentina, and the Université de Lille, France. JAGC especially thanks the French Ministère de l'Enseignement Supérieur et de la Recherche for his post-doctoral grant at the LASIR laboratory in France, as well as the Dirección de Investigaciónes of the UPTC for financial support provided through Project SGI 2343. VPLV acknowledges the Universidad Antonio Nariño for the financial support provided through Project No. 2016231. GAE and OEP are Research Fellows of CONICET.

Funding information

Funding for this research was provided by: the Dirección de Investigaciones de la UPTC (contract No. SGI2343); Universidad Antonio Nariño (contract No. 2016231); CONICET (contract No. (PIP 11220130100651CO)); and UNLP (grant No. (Grant to Project 11/X857)).

References

First citationAddison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349–1356.  CSD CrossRef Web of Science Google Scholar
First citationBanerjee, S., Dixit, A., Maheswaramma, K. S., Maity, B., Mukherjee, S., Kumar, A., Karande, A. A. & Chakravarty, A. R. (2016). J. Chem. Sci. 128, 165–175.  Web of Science CSD CrossRef CAS Google Scholar
First citationBarwiolek, M., Szczęsny, R. & Szłyk, E. (2016). J. Chem. Sci. 128, 1057–1066.  Web of Science CrossRef CAS Google Scholar
First citationBraymer, J. J., Merrill, N. M. & Lim, M. H. (2012). Inorg. Chim. Acta, 380, 261–268.  Web of Science CSD CrossRef CAS Google Scholar
First citationChaviara, A. Th., Christidis, P. C., Papageorgiou, A., Chrysogelou, E., Hadjipavlou-Litina, D. J. & Bolos, C. A. (2005). J. Inorg. Biochem. 99, 2102–2109.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationDeToma, A. S., Salamekh, S., Ramamoorthy, A. & Lim, M. H. (2012). Chem. Soc. Rev. 41, 608–621.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDias, P. M., Kinouti, L., Constantino, V. R. L., Ferreira, A. M. D. C., Gonçalves, M. B., Nascimento, R. R., Petrilli, H. M., Caldas, M. & Frem, R. C. G. (2010). Quím. Nova, 33, 2135–2142.  Web of Science CrossRef CAS Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS 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 citationGupta, K. C. & Sutar, A. K. (2008). Coord. Chem. Rev. 252, 1420–1450.  Web of Science CrossRef CAS Google Scholar
First citationHöfer, P., Grupp, A., Nebenführ, H. & Mehring, M. (1986). Chem. Phys. Lett. 132, 279–282.  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. In the press. https://doi.org/10.1107/S1600576719014092  Google Scholar
First citationMcKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. pp. 3814–3816.  Web of Science CrossRef Google Scholar
First citationRigaku OD (2015). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.  Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). 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 citationTrudu, F., Amato, F., Vaňhara, P., Pivetta, T., Peña-Méndez, E. M. & Havel, J. (2015). J. Appl. Biomed. 13, 79–103.  Web of Science 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. University of Western Australia.  Google Scholar
First citationXavier, V. S. & Murugesan, A. R. (1998). Mol. Phys. 94, 269–273.  CrossRef 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