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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Bis[S-benzyl 3-(furan-2-yl­methyl­­idene)di­thio­carbazato-κ2N3,S]copper(II): crystal structure and Hirshfeld surface analysis

CROSSMARK_Color_square_no_text.svg

aDepartment of Chemistry, Faculty of Science, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor Darul Ehsan, Malaysia, bDiscipline of Chemistry, School of Environmental and Life Sciences, University of Newcastle, University Drive, Callaghan, NSW 2308, Australia, cDepartment of Chemistry, St. Francis Xavier University, PO Box 5000, Antigonish, NS B2G 2W5, Canada, dMaterials Synthesis and Characterization Laboratory, Institute of Advanced Technology, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor Darul Ehsan, Malaysia, eDepartment of Physics, Bhavan's Sheth R. A. College of Science, Ahmedabad, Gujarat 380001, India, and fResearch Centre for Crystalline Materials, School of Science and Technology, Sunway University, 47500 Bandar Sunway, Selangor Darul Ehsan, Malaysia
*Correspondence e-mail: edwardt@sunway.edu.my

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 29 April 2019; accepted 2 May 2019; online 14 May 2019)

The title CuII complex, [Cu(C13H11N2OS2)2], features a trans-N2S2 donor set as a result of the CuII atom being located on a crystallographic centre of inversion and being coordinated by thiol­ate-S and imine-N atoms derived from two di­thio­carbazate anions. The resulting geometry is distorted square-planar. In the crystal, π(chelate ring)–π(fur­yl) [inter-centroid separation = 3.6950 (14) Å and angle of inclination = 5.33 (13)°] and phenyl-C—H⋯π(phen­yl) inter­actions sustain supra­molecular layers lying parallel to ([\overline{1}]02). The most prominent inter­actions between layers, as confirmed by an analysis of the calculated Hirshfeld surface, are phenyl-H⋯H(phen­yl) contacts. Indications for Cu⋯Cg(fur­yl) contacts (Cu⋯Cg = 3.74 Å) were also found. Inter­action energy calculations suggest the contacts between mol­ecules are largely dispersive in nature.

1. Chemical context

Di­thio­carbaza­tes, derived from sulfur–nitro­gen donor ligands were first reviewed in the 1970s (Ali & Livingstone, 1974[Ali, M. A. & Livingstone, S. E. (1974). Coord. Chem. Rev. 13, 101-132.]). These Schiff base mol­ecules are readily prepared from the reaction of primary amines with aldehydes or ketones and are potentially multidentate ligands for metals (Ali et al., 2005[Ali, M. A., Mirza, A. H., Fereday, R. J., Butcher, R. J., Fuller, J. M., Drew, S. C., Gahan, L. R., Hanson, G. R., Moubaraki, B. & Murray, K. S. (2005). Inorg. Chim. Acta, 358, 3937-3948.]; Mokhtaruddin et al., 2017[Mokhtaruddin, N. S. M., Yusof, E. N. M., Ravoof, T. B. S. A., Tiekink, E. R. T., Veerakumarasivam, A. & Tahir, M. I. M. (2017). J. Mol. Struct. 1139, 1-9.]). Schiff bases display significant biological and pharmacological activities that can be tuned by incorporating different types of substituents through the condensation reaction (How et al., 2008[How, F. N. F., Crouse, K. A., Tahir, M. I. M., Tarafder, M. T. H. & Crowley, A. R. (2008). Polyhedron 27, 332-3329.]; Low et al., 2016[Low, M. L., Maigre, L. M., Tahir, M. I. M. T., Tiekink, E. R. T., Dorlet, P., Guillot, R., Ravoof, T. B., Rosli, R., Pagès, J.-M., Policar, C., Delsuc, N. & Crouse, K. A. (2016). Eur. J. Med. Chem. 120, 1-12.]). Transition-metal complexes containing Schiff base ligands have also been intensively studied because of their simple routes of synthesis, the variety of their structural geometries and, particularly pertinent, as small chemical changes often produce wide variations in their bioactivities (Mirza et al., 2014[Mirza, A. H., Hamid, M. H. S. A., Aripin, S., Karim, M. R., Arifuzzaman, M., Ali, M. A. & Bernhardt, P. V. (2014). Polyhedron, 74, 16-23.]; Zangrando et al., 2015[Zangrando, E., Islam, M. T., Islam, M. A. A. A., Sheikh, M. C., Tarafder, M. T. H., Miyatake, R., Zahan, R. & Hossain, M. A. (2015). Inorg. Chim. Acta, 427, 278-284.]; Lima et al., 2018[Lima, F. C., Silva, T. S., Martins, C. H. G. & Gatto, C. C. (2018). Inorg. Chim. Acta, 483, 464-472.]). Recently, a copper(II) di­thio­carbazate complex containing a Schiff base derived from S-hexyl­dithio­carbazate and 4-methyl­benzaldehyde was reported to have excellent anti-bacterial activity against Escherichia coli (Zangrando et al., 2017[Zangrando, E., Begum, M. S., Sheikh, M. C., Miyatake, R., Hossain, M. M., Alam, M. M., Hasnat, M. A., Halim, M. A., Ahmed, S., Rahman, M. N. & Ghosh, A. (2017). Arabian J. Chem. 10, 172-184.]). More recently, investigators have reported the potent biological activity of a copper(II) complex that contained a tridentate Schiff base derived from S-benzyl­dithio­carbazate and 2-hy­droxy-5-(phenyl­diazen­yl)benzaldehyde against a human cervical cancer line (HeLa) (Kongot et al., 2019[Kongot, M., Reddy, D., Singh, V., Patel, R., Singhal, N. K. & Kumar, A. (2019). Spectrochim. Acta A 212, 330-342.]). The copper(II) complex had comparable biological activities as the well-known anti-cancer drug cisplatin against the tested cells (Kongot et al., 2019[Kongot, M., Reddy, D., Singh, V., Patel, R., Singhal, N. K. & Kumar, A. (2019). Spectrochim. Acta A 212, 330-342.]). As part of on-going studies in the structural chemistry and potential bioactivity of copper(II) complexes containing di­thio­carbazate Schiff base ligands, herein the synthesis of the title copper(II) complex, (I)[link], its single crystal X-ray diffraction analysis and a detailed study of supra­molecular association by an analysis of calculated Hirshfeld surfaces and computation chemistry are described.

[Scheme 1]

2. Structural commentary

The mol­ecular structure of (I)[link], Fig. 1[link], has the CuII atom located on a crystallographic centre of inversion and coordin­ated by two chelating di­thio­carbazate anions, each via the thiol­ate-S and imine-N atoms (Table 1[link]). The resulting trans-N2S2 donor set defines a distorted square-planar geometry: the major distortion from the ideal angles subtended at the copper atom is the acute S1—Cu—N2 chelate angle of 85.83 (6)°. The conformation about the endocylic imine bond is Z, as a result of chelation, whereas the exocyclic imine bond has an E conformation.

Table 1
Selected geometric parameters (Å, °)

Cu—S1 2.1845 (7) N1—N2 1.409 (3)
Cu—N2 1.923 (2) C1—N1 1.286 (3)
C1—S1 1.720 (3) C9—N2 1.300 (3)
C1—S2 1.753 (2)    
       
S1—Cu—N2 85.83 (6) S1—C1—N1 125.08 (19)
S1i—Cu—N2 94.18 (6) S2—C1—N1 119.9 (2)
S1—C1—S2 115.03 (15)    
Symmetry code: (i) -x+1, -y+1, -z+1.
[Figure 1]
Figure 1
The mol­ecular structure of (I)[link] showing the atom-labelling scheme and displacement ellipsoids at the 70% probability level. Unlabelled atoms are related by the symmetry operation 1 − x, 1 − y, 1 − z.

The bidentate mode of the coordination of the di­thio­carbazate ligand leads to the formation of five-membered CuN2CS chelate rings. While the r.m.s. deviation for the five atoms is relatively small at 0.0453 Å, suggesting a near planar ring, a better description for the conformation is that of an envelope with the copper atom being the flap atom. In this description, the r.m.s. deviation of the S1, N1, N2 and N3 atoms of the ring is 0.0002 Å, with the Cu atom lying 0.199 (3) Å out of the plane. The dihedral angle between the best plane through the chelate ring and the 2-furyl ring is 5.33 (18)° indicating an essentially co-planar relationship. By contrast, the dihedral between the chelate and phenyl rings is 86.75 (7)°, indicative of an orthogonal relationship. Finally, the dihedral angle between the peripheral organic rings is 81.42 (9)°.

The structure of the acid form of the anion in (I)[link] is available for comparison (Shan et al., 2008[Shan, S., Tian, Y.-L., Wang, S.-H., Wang, W.-L. & Xu, Y.-L. (2008). Acta Cryst. E64, o1024.]). Referring to the data in Table 1[link], significant changes in key bond lengths have occurred upon deprotonation and coordination of the mol­ecule to CuII in (I)[link]. Thus, the C1—S1 [1.669 (2) Å for the acid], N1—N2 [1.381 (2) Å] and C9—N2 [1.280 (3) Å] bond lengths have all elongated in (I)[link], Table 1[link], while the C1—N1 bond length has shortened [1.336 (3) Å]. Significant changes in the angles subtended at the quaternary C1 atom are also noted, in particular for the S1—C1—S2 angle which has narrowed by ca 10° in (I)[link] from 124.76 (12)° in the acid with concomitant widening of the S2—C1–N1 angle by ca 5°, changes consistent with the reorganization of π-electron density from the C1—S1 to C1—N1 bonds in (I)[link].

3. Supra­molecular features

The most prominent feature of the mol­ecular packing is the formation of supra­molecular layers lying parallel to ([\overline{1}]02), Fig. 2[link](a). The association between mol­ecules is of the type π(chelate ring)–π(fur­yl) whereby the inter-centroid Cg(Cu,S1,N1,N2,C1)–Cg(O1,C10–C13)i separation is 3.6950 (14) Å with angle of inclination = 5.33 (13)°; symmetry operation (i) x, −1 + y, z. Such ππ inter­actions between chelate rings and aromatic rings are well documented in the literature, especially for sterically unencumbered square-planar complexes and can impart significant energies of stabilization to the mol­ecular packing (Malenov et al. 2017[Malenov, D. P., Janjić, G. V., Medaković, V. B., Hall, M. B. & Zarić, S. D. (2017). Coord. Chem. Rev. 345, 318-341.]; Tiekink, 2017[Tiekink, E. R. T. (2017). Coord. Chem. Rev. 345, 209-228.]). In the present case, these inter­actions link mol­ecules along the b-axis direction. Links between the chains to form layers are of the type phenyl-C—H⋯π(phen­yl), Table 2[link]. A view of the unit-cell contents is shown in Fig. 2[link](b). Details of the weak inter­molecular contacts connecting layers are given in the analysis of the calculated Hirshfeld surfaces below.

Table 2
Hydrogen-bond geometry (Å, °)

Cg1 is the centroid of the (C3–C8) ring.

D—H⋯A D—H H⋯A DA D—H⋯A
C5—H5⋯Cg1ii 0.95 2.96 3.646 (3) 131
Symmetry code: (ii) [-x, y+{\script{1\over 2}}, -z+{\script{1\over 2}}].
[Figure 2]
Figure 2
Mol­ecular packing in (I)[link]: (a) a view of the supra­molecular layer sustained by π(chelate ring)–π(fur­yl) and phenyl-C—H⋯π(phen­yl) inter­actions shown as blue and purple dashed lines, respectively, and (b) a view of the unit-cell contents shown in projection down the b axis highlighting the stacking of layers.

4. Analysis of the Hirshfeld surfaces

The analysis of the Hirshfeld surfaces calculated for (I)[link] was conducted as per literature precedents (Tan et al., 2019[Tan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308-318.]) employing Crystal Explorer (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 v17. The University of Western Australia.]). The assumption of the inter­molecular C—H⋯π contact in the crystal of (I)[link] is justified through the diminutive red spots near the phenyl-C4 and H5 atoms on the Hirshfeld surfaces mapped over dnorm in Fig. 3[link]. The short inter­atomic H⋯H contact, involving phenyl H8 atoms and occurring between layers, and the C⋯C contact, between the methyl­ene-C9 and furyl-C11 atoms, are also evident as the faint-red spots near the respective atoms in Fig. 3[link]. On the Hirshfeld surfaces mapped over electrostatic potential in Fig. 4[link], the donors and acceptors of inter­molecular C—H⋯π contacts, Table 2[link], are viewed as blue bumps and light-red concave regions, respectively. Also, the short inter­atomic S⋯H/H⋯S contacts, which are electrostatic in nature, Table 3[link], show red and blue regions about the respective atoms. The environment around a reference mol­ecule within the Hirshfeld surface mapped with the shape-index property is illustrated in Fig. 5[link], and highlights the C—H⋯π/π⋯H—C contacts.

Table 3
Summary of short inter­atomic contacts (Å) in (I)

Contact Distance Symmetry operation
H8⋯H8 2.11 x, 1 − y, 1 − z
H5⋯C4 2.66 x, [{1\over 2}] + y, [{1\over 2}] − z
H2B⋯S2 2.97 x, 1 + y, z
C9⋯C11 3.364 (4) x, −1 + y, z
Notes: (a) The inter­atomic distances are calculated in Crystal Explorer (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 v17. The University of Western Australia.]) whereby the X—H bond lengths are adjusted to their neutron values.
[Figure 3]
Figure 3
A view of the Hirshfeld surface for (I)[link] mapped over dnorm in the range −0.080 to +1.213 arbitrary units.
[Figure 4]
Figure 4
A view of the Hirshfeld surface for (I)[link] mapped over the electrostatic potential in the range −0.036 to + 0.034 atomic units.
[Figure 5]
Figure 5
A view of the Hirshfeld surface with the shape-index property highlighting C—H⋯π/π⋯H—C contacts by black dotted lines.

The overall two-dimensional fingerprint plot, Fig. 6[link](a), and those delineated into H⋯H, C⋯H/H⋯C, S⋯H/H⋯S and C⋯C contacts are illustrated in Fig. 6[link](b)–(e), respectively; the percentage contribution from all the identified inter­atomic contacts to the Hirshfeld surface are summarized qu­anti­tatively in Table 4[link].

Table 4
Percentage contributions of inter­atomic contacts to the Hirshfeld surface for (I)

Contact Percentage contribution
H⋯H 36.2
C⋯H/H⋯C 23.0
S⋯H/H⋯S 17.5
O⋯H/H⋯O 5.1
C⋯N/N⋯C 3.3
S⋯O/O⋯S 2.9
N⋯H/H⋯N 2.8
Cu⋯C/C⋯Cu 2.7
C⋯C 2.6
C⋯S/S⋯C 1.3
N⋯S/S⋯N 1.2
O⋯O 0.5
N⋯O/O⋯N 0.3
N⋯N 0.3
Cu⋯N/N⋯Cu 0.2
Cu⋯H/H⋯Cu 0.1
Cu⋯O/O⋯Cu 0.1
[Figure 6]
Figure 6
(a) The full two-dimensional fingerprint plot for (I)[link] and fingerprint plots delineated into (b) H⋯H, (c) C⋯H/H⋯C, (d) S⋯H/H⋯S and (e) C⋯C contacts.

The conical tip appearing at de + di ∼2.1 Å in the fingerprint plot delineated into H⋯H contacts in Fig. 6[link](b), represents the short inter-layer H⋯H contact involving phenyl-H8 atoms, Table 3[link]. The presence of the C—H⋯π inter­action is evident through the short inter­atomic C⋯H/H⋯C contact characterized as the pair of forceps-like tips at de + di ∼2.7 Å in the respective delineated fingerprint plot of Fig. 6[link](c) and Table 3[link]. In the fingerprint plot delineated into S⋯H/H⋯S contacts, Fig. 6[link](d), the short inter­atomic contact involving the S-benzyl atoms, Table 3[link], appear as the pair of forceps-like tips at de + di < 3.0 Å, i.e. at the sum of van der Waals radii. The distribution of points in the fingerprint plot delineated into C⋯C contacts, Fig. 6[link](e), forming triangular tip at de + di ∼3.3 Å is due to the presence of such short inter­atomic contacts summarized in Table 3[link]. The presence of inter­molecular ππ stacking between chelate and furyl rings results in the small but significant percentage contribution from the participating atoms, as listed in Table 4[link]. The small contributions from the other remaining inter­atomic contacts summarized in Table 4[link] have a negligible effect on the packing.

5. Computational chemistry

Utilizing Crystal Explorer (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 v17. The University of Western Australia.]), the pairwise inter­action energies between the mol­ecules within the crystal were calculated by summing up four energy component, namely electrostatic (Eele), polarization (Epol), dispersion (Edis) and exchange-repulsion (Erep). The energies were obtained using the wave function calculated at the HF/STO-3G level theory. The strength and nature of the inter­molecular inter­actions are summarized qu­anti­tatively in Table 5[link]. From the inter­action energies calculated between the reference mol­ecule and the symmetry-related mol­ecule at x, −1 + y, z in Table 5[link], it is observed that the greatest energy value is due to the combined influence of Cu⋯furyl [Cu⋯Cg(fur­yl) = 3.74 Å], π(chelate)–π(fur­yl), C⋯C and S⋯H/H⋯S inter­actions. Among these inter­actions, the short inter­atomic S⋯H/H⋯S contact contributes to the electrostatic component while the others to the dispersion component of the energies. Even though the inter-centroid distance between symmetry-related phenyl (C3–C8) rings are greater than 4.0 Å [CgCgi = 4.3102 (17) Å; (i) − x, 2 − y, 1 − z] and the inter­atomic S⋯H distance is greater than sum of their van der Waal radii (S1⋯H11ii = 3.11 Å; x, [{3\over 2}] − y, −[{1\over 2}] + z), they possess greater inter­action energies compared to inter­molecular phenyl-C—H⋯π(phen­yl) and short inter­atomic H⋯H contacts, as summarized in Table 5[link]. The magnitudes of the inter­molecular energies are represented graphically in the energy frameworks down the b-axis direction in Fig. 7[link]. Here, the supra­molecular architecture of crystals is viewed through the cylinders joining the centroids of mol­ecular pairs by using red, green and blue colour codes for the components Eele, Edisp and Etot, respectively; the radius of the cylinder is proportional to the magnitude of inter­action energy. It is clearly evident from the energy frameworks shown in Fig. 7[link] that the major contribution to the inter­molecular inter­actions is from the dispersion energy component in the absence of conventional hydrogen bonds in the crystal.

Table 5
Summary of inter­action energies (kJ mol−1) calculated for (I)

Contact R (Å) Eele Epol Edis Erep Etot
Cu⋯Cg(fur­yl)i +            
Cg(chelate)⋯Cg(fur­yl)i +            
C9⋯C11i +            
S2⋯H2Bi 5.02 −23.2 −9.4 −154.4 97.6 −89.7
Cg(phen­yl)⋯Cg(phen­yl)ii 16.15 −6.3 −3.3 −50.9 28.3 −31.5
S1⋯H11iii 11.25 −12.0 −2.6 −10.6 5.2 −19.2
C5—H5⋯Cg(phen­yl)iv 17.06 −6.2 −2.1 −20.6 13.8 −15.1
H8⋯H8v 15.35 0.7 −0.9 −15.6 7.9 −7.5
Notes: Symmetry operations: (i) x, −1 + y, z; (ii) −x, 2 − y, 1 − z; (iii) x, [{3\over 2}] − y, [{1\over 2}] + z; (iv) −x, [{1\over 2}] + y, [{1\over 2}] − z; (v) −x, 1 − y, 1 − z.
[Figure 7]
Figure 7
The energy frameworks viewed down the b-axis direction comprising (a) electrostatic potential force, (b) dispersion force and (c) total energy for a cluster about a reference mol­ecule of (I)[link]. The energy frameworks were adjusted to the same scale factor of 50 with a cut-off value of 3 kJ mol−1 within 2 × 2 × 2 unit cells.

6. Database survey

The Cambridge Structural Database (Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) contains just about 100 structures with the basic core found in (I)[link]. Manual sorting to identify ligands without additional donors as in (I)[link], e.g. substituents carrying pyridyl or phenoxide, neutral mol­ecules only and non-solvated structures yielded 24 analogues to (I)[link] with deposited atomic coordinates. Eleven of these structures adopt the trans-N2S2 square-planar geom­etry as in (I)[link], while the remaining 13 structures adopt a flattened tetra­hedral coordination geometry. The structural di­versity exhibited by these complexes is emphasized by the bi­nuclear species [Cu{SCS[(CH2)5Me]=NN=CC6H4OMe-4}2]2 arising from inter­molecular Cu⋯S inter­actions between centrosymmetrically related trans-N2S2 square-planar geometries (Begum et al., 2017[Begum, M. S., Zangrando, E., Sheikh, M. C., Miyatake, R., Howlader, M. B. H., Rahman, M. N. & Ghosh, A. (2017). Transit. Met. Chem. 42, 553-563.]).

7. Synthesis and crystallization

Synthesis of the 2-furaldehyde Schiff base of S-benzyl­dithio­carbazate: S-Benzyl­dithio­carbazate (SBDTC) was synthesized following a procedure adapted from a previous report (Tarafder et al., 2001[Tarafder, M. T. H., Kasbollah, A., Crouse, K. A., Ali, A. M., Yamin, B. M. & Fun, H.-K. (2001). Polyhedron, 20, 2363-2370.]). The Schiff base was synthesized using a procedure adapted from the literature (Yusof et al., 2015[Yusof, E. N. M., Ravoof, T. B. S. A., Tiekink, E. R. T., Veerakumarasivam, A., Crouse, K. A., Tahir, M. I. M. & Ahmad, H. (2015). Int. J. Mol. Sci. 16, 11034-11054.]) by reacting SBDTC (3.96 g, 0.02 mol) and an equimolar amount of 2-furaldehyde (1.92 g, 0.02 mmol) in hot ethanol (20 ml). The mixture was then heated until the volume reduced to half, followed by stirring under room temperature until a precipitate had formed. The resulting Schiff base was then washed with ice-cold ethanol, recrystallized from ethanol solution and dried over silica gel. Colour: Yellow. Yield 94%, m.p. 447–449 K. Elemental analysis: Calculated for C13H12N2OS2: C, 56.49; H, 4.38; N, 10.14. Found; C, 56.64; H, 4.21; N, 9.64. FTIR (ATR, cm−1): 3089 (w) ν(N—H), 1609 (m) ν(C=N), 1016 (s) ν(N—N), 763 (s), ν(C=S).

Synthesis of (I)[link]: The Schiff base synthesized above (0.55 g, 0.002 mol) was dissolved in hot ethanol (50 ml) and added to copper(II) acetate monohydrate (0.20 g, 0.001 mol) in an ethano­lic solution (30 ml). The mixture was heated until the volume of the solution reduced to half. Precipitation occurred once the mixture had cooled to room temperature. The precipitate was filtered and dried over silica gel. The title complex was recrystallized from methanol solution as dark-brown prisms in 91% yield. M.p. 456–458 K. Elemental analysis: Calculated for C26H22CuN4O2S4: C, 50.84; H, 3.61; N, 9.12; Cu, 10.34. Found; C, 50.49; H, 3.45; N, 8.77; Cu, 10.81. FTIR (ATR, cm−1): 1593 (m), ν(C=N), 964 (s), ν(N—N), 760 (s), ν(C—S).

8. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 6[link]. The carbon-bound H atoms were placed in calculated positions (C—H = 0.95–0.99 Å) and were included in the refinement in the riding-model approximation, with Uiso(H) set to 1.2Ueq(C).

Table 6
Experimental details

Crystal data
Chemical formula [Cu(C13H11N2OS2)2]
Mr 614.25
Crystal system, space group Monoclinic, P21/c
Temperature (K) 100
a, b, c (Å) 15.3515 (7), 5.0151 (3), 16.7186 (8)
β (°) 94.618 (4)
V3) 1282.98 (11)
Z 2
Radiation type Mo Kα
μ (mm−1) 1.21
Crystal size (mm) 0.30 × 0.20 × 0.10
 
Data collection
Diffractometer Agilent Xcalibur Eos Gemini
Absorption correction Multi-scan (CrysAlis PRO; Agilent, 2011[Agilent (2011). CrysAlis PRO. Agilent Technologies, Yarnton, England.])
Tmin, Tmax 0.744, 1.000
No. of measured, independent and observed [I > 2σ(I)] reflections 5864, 2898, 2382
Rint 0.027
(sin θ/λ)max−1) 0.677
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.040, 0.111, 1.04
No. of reflections 2898
No. of parameters 169
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.49, −0.61
Computer programs: CrysAlis PRO (Agilent, 2011[Agilent (2011). CrysAlis PRO. Agilent Technologies, Yarnton, England.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2014 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), ORTEP-3 for Windows (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]), DIAMOND (Brandenburg, 2006[Brandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

Data collection: CrysAlis PRO (Agilent, 2011); cell refinement: CrysAlis PRO (Agilent, 2011); data reduction: CrysAlis PRO (Agilent, 2011); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008); program(s) used to refine structure: SHELXL2014 (Sheldrick, 2015); molecular graphics: ORTEP-3 for Windows (Farrugia, 2012) and DIAMOND (Brandenburg, 2006); software used to prepare material for publication: publCIF (Westrip, 2010).

Bis[S-benzyl 3-(furan-2-ylmethylidene)dithiocarbazato-κ2N3,S]copper(II) top
Crystal data top
[Cu(C13H11N2OS2)2]F(000) = 630
Mr = 614.25Dx = 1.590 Mg m3
Monoclinic, P21/cMo Kα radiation, λ = 0.71073 Å
a = 15.3515 (7) ÅCell parameters from 1956 reflections
b = 5.0151 (3) Åθ = 2.4–28.7°
c = 16.7186 (8) ŵ = 1.21 mm1
β = 94.618 (4)°T = 100 K
V = 1282.98 (11) Å3Prism, dark-brown
Z = 20.30 × 0.20 × 0.10 mm
Data collection top
Agilent Xcalibur Eos Gemini
diffractometer
2898 independent reflections
Radiation source: fine-focus sealed X-ray tube, Enhance (Mo) X-ray Source2382 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.027
Detector resolution: 16.1952 pixels mm-1θmax = 28.8°, θmin = 2.4°
ω scansh = 1918
Absorption correction: multi-scan
(CrysAlis PRO; Agilent, 2011)
k = 56
Tmin = 0.744, Tmax = 1.000l = 2220
5864 measured reflections
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.040Hydrogen site location: inferred from neighbouring sites
wR(F2) = 0.111H-atom parameters constrained
S = 1.03 w = 1/[σ2(Fo2) + (0.0542P)2 + 1.1451P]
where P = (Fo2 + 2Fc2)/3
2898 reflections(Δ/σ)max < 0.001
169 parametersΔρmax = 0.49 e Å3
0 restraintsΔρmin = 0.61 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
Cu0.50000.50000.50000.01942 (15)
S10.40280 (4)0.46830 (15)0.39759 (4)0.02392 (18)
S20.22892 (4)0.71156 (14)0.37681 (4)0.02121 (18)
O10.42829 (12)1.2573 (4)0.69781 (11)0.0240 (4)
N10.34226 (13)0.8207 (4)0.49969 (13)0.0186 (5)
N20.42578 (13)0.7729 (4)0.53831 (13)0.0168 (5)
C10.32893 (16)0.6845 (5)0.43470 (16)0.0185 (5)
C20.17122 (17)0.9561 (6)0.43348 (17)0.0221 (6)
H2A0.17560.90750.49110.027*
H2B0.19761.13470.42810.027*
C30.07652 (17)0.9609 (5)0.40095 (16)0.0192 (6)
C40.04586 (19)1.1481 (6)0.34526 (17)0.0255 (6)
H40.08481.27740.32680.031*
C50.0419 (2)1.1495 (6)0.31570 (18)0.0299 (7)
H50.06231.28000.27750.036*
C60.09871 (18)0.9637 (6)0.34150 (17)0.0236 (6)
H60.15830.96470.32120.028*
C70.06860 (19)0.7743 (6)0.39746 (19)0.0301 (7)
H70.10760.64530.41600.036*
C80.01881 (19)0.7741 (6)0.42627 (19)0.0321 (7)
H80.03930.64260.46420.039*
C90.44368 (17)0.9293 (6)0.59946 (15)0.0192 (5)
H90.49910.90360.62800.023*
C100.39017 (17)1.1365 (5)0.62925 (15)0.0188 (5)
C110.37208 (19)1.4495 (6)0.71860 (17)0.0250 (6)
H110.38191.56410.76370.030*
C120.30047 (19)1.4555 (6)0.66655 (17)0.0246 (6)
H120.25181.57150.66850.030*
C130.31160 (18)1.2562 (5)0.60850 (17)0.0224 (6)
H130.27211.21370.56370.027*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Cu0.0141 (2)0.0252 (3)0.0192 (3)0.00134 (18)0.00266 (17)0.00100 (19)
S10.0163 (3)0.0344 (4)0.0207 (3)0.0034 (3)0.0001 (3)0.0076 (3)
S20.0153 (3)0.0256 (4)0.0222 (3)0.0000 (3)0.0016 (2)0.0025 (3)
O10.0217 (10)0.0322 (11)0.0178 (9)0.0049 (8)0.0005 (7)0.0047 (8)
N10.0126 (10)0.0229 (12)0.0202 (11)0.0010 (9)0.0013 (8)0.0005 (9)
N20.0106 (10)0.0216 (11)0.0186 (10)0.0031 (8)0.0027 (8)0.0015 (9)
C10.0140 (12)0.0192 (13)0.0224 (13)0.0040 (10)0.0020 (10)0.0024 (11)
C20.0168 (13)0.0239 (14)0.0251 (14)0.0007 (11)0.0019 (10)0.0021 (11)
C30.0163 (12)0.0223 (14)0.0187 (13)0.0023 (10)0.0002 (10)0.0046 (11)
C40.0253 (14)0.0257 (15)0.0251 (14)0.0018 (12)0.0008 (11)0.0018 (12)
C50.0298 (16)0.0312 (16)0.0276 (15)0.0017 (13)0.0052 (12)0.0073 (13)
C60.0195 (13)0.0282 (15)0.0225 (14)0.0060 (11)0.0018 (11)0.0057 (12)
C70.0193 (14)0.0341 (17)0.0366 (17)0.0021 (12)0.0012 (12)0.0086 (14)
C80.0232 (14)0.0346 (17)0.0376 (17)0.0006 (13)0.0030 (12)0.0166 (14)
C90.0137 (12)0.0264 (14)0.0176 (12)0.0014 (10)0.0018 (10)0.0006 (11)
C100.0184 (12)0.0225 (14)0.0158 (12)0.0050 (11)0.0032 (10)0.0014 (11)
C110.0291 (15)0.0280 (15)0.0184 (13)0.0017 (12)0.0052 (11)0.0007 (12)
C120.0231 (14)0.0230 (14)0.0280 (15)0.0028 (11)0.0045 (11)0.0004 (12)
C130.0211 (13)0.0213 (14)0.0245 (14)0.0000 (11)0.0007 (11)0.0017 (11)
Geometric parameters (Å, º) top
Cu—S12.1845 (7)C4—C51.397 (4)
Cu—N21.923 (2)C4—H40.9500
Cu—N2i1.923 (2)C5—C61.369 (4)
Cu—S1i2.1845 (7)C5—H50.9500
C1—S11.720 (3)C6—C71.386 (4)
C1—S21.753 (2)C6—H60.9500
S2—C21.823 (3)C7—C81.389 (4)
O1—C111.358 (3)C7—H70.9500
O1—C101.384 (3)C8—H80.9500
N1—N21.409 (3)C9—C101.438 (4)
C1—N11.286 (3)C9—H90.9500
C9—N21.300 (3)C10—C131.367 (4)
C2—C31.511 (3)C11—C121.346 (4)
C2—H2A0.9900C11—H110.9500
C2—H2B0.9900C12—C131.413 (4)
C3—C81.379 (4)C12—H120.9500
C3—C41.378 (4)C13—H130.9500
N2—Cu—N2i180.00 (11)C6—C5—C4120.4 (3)
S1—Cu—N285.83 (6)C6—C5—H5119.8
N2i—Cu—S194.17 (6)C4—C5—H5119.8
S1i—Cu—N294.18 (6)C5—C6—C7119.5 (3)
N2i—Cu—S1i85.82 (6)C5—C6—H6120.2
S1—Cu—S1i180.0C7—C6—H6120.2
C1—S1—Cu95.74 (9)C6—C7—C8119.7 (3)
C1—S2—C2101.88 (12)C6—C7—H7120.2
C11—O1—C10106.7 (2)C8—C7—H7120.2
C1—N1—N2112.0 (2)C3—C8—C7121.3 (3)
C9—N2—N1112.6 (2)C3—C8—H8119.4
C9—N2—Cu126.72 (18)C7—C8—H8119.4
N1—N2—Cu120.67 (16)N2—C9—C10128.2 (2)
S1—C1—S2115.03 (15)N2—C9—H9115.9
S1—C1—N1125.08 (19)C10—C9—H9115.9
S2—C1—N1119.9 (2)C13—C10—O1108.9 (2)
C3—C2—S2108.50 (18)C13—C10—C9138.3 (2)
C3—C2—H2A110.0O1—C10—C9112.8 (2)
S2—C2—H2A110.0C12—C11—O1110.7 (2)
C3—C2—H2B110.0C12—C11—H11124.7
S2—C2—H2B110.0O1—C11—H11124.7
H2A—C2—H2B108.4C11—C12—C13106.9 (3)
C8—C3—C4118.5 (3)C11—C12—H12126.6
C8—C3—C2120.1 (2)C13—C12—H12126.6
C4—C3—C2121.4 (3)C10—C13—C12106.8 (2)
C3—C4—C5120.6 (3)C10—C13—H13126.6
C3—C4—H4119.7C12—C13—H13126.6
C5—C4—H4119.7
C1—N1—N2—C9173.3 (2)C5—C6—C7—C80.4 (5)
C1—N1—N2—Cu6.9 (3)C4—C3—C8—C70.7 (5)
N2—N1—C1—S10.0 (3)C2—C3—C8—C7179.9 (3)
N2—N1—C1—S2179.76 (17)C6—C7—C8—C30.7 (5)
Cu—S1—C1—N15.3 (2)N1—N2—C9—C100.8 (4)
Cu—S1—C1—S2174.52 (13)Cu—N2—C9—C10179.4 (2)
C2—S2—C1—N11.1 (3)C11—O1—C10—C130.4 (3)
C2—S2—C1—S1179.07 (15)C11—O1—C10—C9179.1 (2)
C1—S2—C2—C3168.47 (19)N2—C9—C10—C134.3 (6)
S2—C2—C3—C882.6 (3)N2—C9—C10—O1177.5 (2)
S2—C2—C3—C496.8 (3)C10—O1—C11—C120.0 (3)
C8—C3—C4—C50.5 (4)O1—C11—C12—C130.4 (3)
C2—C3—C4—C5179.9 (3)O1—C10—C13—C120.6 (3)
C3—C4—C5—C60.2 (5)C9—C10—C13—C12178.8 (3)
C4—C5—C6—C70.2 (5)C11—C12—C13—C100.6 (3)
Symmetry code: (i) x+1, y+1, z+1.
Hydrogen-bond geometry (Å, º) top
Cg1 is the centroid of the (C3–C8) ring.
D—H···AD—HH···AD···AD—H···A
C5—H5···Cg1ii0.952.963.646 (3)131
Symmetry code: (ii) x, y+1/2, z+1/2.
Summary of short interatomic contacts (Å) in (I) top
ContactDistanceSymmetry operation
H8···H82.11-x, 1 - y, 1 - z
H5···C42.66-x, 1/2 + y, 1/2 - z
H2B···S22.97x, 1 + y, z
C9···C113.364 (4)x, -1 + y, z
Notes: (a) The interatomic distances are calculated in Crystal Explorer (Turner et al., 2017) whereby the X—H bond lengths are adjusted to their neutron values.
Percentage contributions of interatomic contacts to the Hirshfeld surface for (I) top
ContactPercentage contribution
H···H36.2
C···H/H···C23.0
S···H/H···S17.5
O···H/H···O5.1
C···N/N···C3.3
S···O/O···S2.9
N···H/H···N2.8
Cu···C/C···Cu2.7
C···C2.6
C···S/S···C1.3
N···S/S···N1.2
O···O0.5
N···O/O···N0.3
N···N0.3
Cu···N/N···Cu0.2
Cu···H/H···Cu0.1
Cu···O/O···Cu0.1
Summary of interaction energies (kJ mol-1) calculated for (I) top
ContactR (Å)EeleEpolEdisErepEtot
Cu···Cg(furyl)i +
Cg(chelate)···Cg(furyl)i +
C9···C11i +
S2···H2Bi5.02-23.2-9.4-154.497.6-89.7
Cg(phenyl)···Cg(phenyl)ii16.15-6.3-3.3-50.928.3-31.5
S1···H11iii11.25-12.0-2.6-10.65.2-19.2
C5—H5···Cg(phenyl)iv17.06-6.2-2.1-20.613.8-15.1
H8···H8v15.350.7-0.9-15.67.9-7.5
Notes: Symmetry operations: (i) x, -1 + y, z; (ii) -x, 2 - y, 1 - z; (iii) x, 3/2 - y, 1/2 + z; (iv) -x, 1/2 + y, 1/2 - z; (v) -x, 1 - y, 1 - z.
 

Footnotes

Additional correspondence author, e-mail: thahira@upm.edu.my.

Acknowledgements

The authors thank the Department of Chemistry, Universiti Putra Malaysia, Malaysia, for facilities to carry out this research. This research was funded by the Universiti Putra Malaysia under the Putra Group Initiative (IPB No. 9581001), Research University Grant Scheme (RUGS No. 9548700) and the Malaysian Fundamental Research Grant Scheme (FRGS No. 01-01-16-1833FR). ENMY wishes to thank the Ministry of Higher Education Malaysia (MoHE) for the award of MyPhD and MyBrain15 scholarships, and also the University of Newcastle for the award of a University of Newcastle Inter­national Postgraduate Research Scholarship (UNIPRS) and a University of Newcastle Research Scholarship Central (UNRSC). NMN wishes to thank MoHE for a MyBrain Science scholarship.

Funding information

Funding for this research was provided by: Universiti Putra Malaysia: Putra Group Initiative (grant No. IPB No. 9581001); Research University Grant Scheme (award No. RUGS No. 9548700); Malaysian Fundamental Research Grant Scheme (grant No. FRGS No. 01-01-16-1833FR).

References

First citationAgilent (2011). CrysAlis PRO. Agilent Technologies, Yarnton, England.  Google Scholar
First citationAli, M. A. & Livingstone, S. E. (1974). Coord. Chem. Rev. 13, 101–132.  CrossRef CAS Google Scholar
First citationAli, M. A., Mirza, A. H., Fereday, R. J., Butcher, R. J., Fuller, J. M., Drew, S. C., Gahan, L. R., Hanson, G. R., Moubaraki, B. & Murray, K. S. (2005). Inorg. Chim. Acta, 358, 3937–3948.  Web of Science CSD CrossRef CAS Google Scholar
First citationBegum, M. S., Zangrando, E., Sheikh, M. C., Miyatake, R., Howlader, M. B. H., Rahman, M. N. & Ghosh, A. (2017). Transit. Met. Chem. 42, 553–563.  Web of Science CSD CrossRef CAS Google Scholar
First citationBrandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.  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 citationHow, F. N. F., Crouse, K. A., Tahir, M. I. M., Tarafder, M. T. H. & Crowley, A. R. (2008). Polyhedron 27, 332–3329.  Web of Science CSD CrossRef Google Scholar
First citationKongot, M., Reddy, D., Singh, V., Patel, R., Singhal, N. K. & Kumar, A. (2019). Spectrochim. Acta A 212, 330–342.  CrossRef CAS Google Scholar
First citationLima, F. C., Silva, T. S., Martins, C. H. G. & Gatto, C. C. (2018). Inorg. Chim. Acta, 483, 464–472.  Web of Science CSD CrossRef CAS Google Scholar
First citationLow, M. L., Maigre, L. M., Tahir, M. I. M. T., Tiekink, E. R. T., Dorlet, P., Guillot, R., Ravoof, T. B., Rosli, R., Pagès, J.-M., Policar, C., Delsuc, N. & Crouse, K. A. (2016). Eur. J. Med. Chem. 120, 1–12.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationMalenov, D. P., Janjić, G. V., Medaković, V. B., Hall, M. B. & Zarić, S. D. (2017). Coord. Chem. Rev. 345, 318–341.  Web of Science CrossRef CAS Google Scholar
First citationMirza, A. H., Hamid, M. H. S. A., Aripin, S., Karim, M. R., Arifuzzaman, M., Ali, M. A. & Bernhardt, P. V. (2014). Polyhedron, 74, 16–23.  Web of Science CSD CrossRef CAS Google Scholar
First citationMokhtaruddin, N. S. M., Yusof, E. N. M., Ravoof, T. B. S. A., Tiekink, E. R. T., Veerakumarasivam, A. & Tahir, M. I. M. (2017). J. Mol. Struct. 1139, 1–9.  Web of Science CSD CrossRef CAS Google Scholar
First citationShan, S., Tian, Y.-L., Wang, S.-H., Wang, W.-L. & Xu, Y.-L. (2008). Acta Cryst. E64, o1024.  Web of Science CSD CrossRef IUCr Journals 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 citationTan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308–318.  Web of Science CrossRef IUCr Journals Google Scholar
First citationTarafder, M. T. H., Kasbollah, A., Crouse, K. A., Ali, A. M., Yamin, B. M. & Fun, H.-K. (2001). Polyhedron, 20, 2363–2370.  Web of Science CSD CrossRef CAS Google Scholar
First citationTiekink, E. R. T. (2017). Coord. Chem. Rev. 345, 209–228.  Web of Science CrossRef CAS 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 v17. The University of Western Australia.  Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationYusof, E. N. M., Ravoof, T. B. S. A., Tiekink, E. R. T., Veerakumarasivam, A., Crouse, K. A., Tahir, M. I. M. & Ahmad, H. (2015). Int. J. Mol. Sci. 16, 11034–11054.  Web of Science CAS PubMed Google Scholar
First citationZangrando, E., Begum, M. S., Sheikh, M. C., Miyatake, R., Hossain, M. M., Alam, M. M., Hasnat, M. A., Halim, M. A., Ahmed, S., Rahman, M. N. & Ghosh, A. (2017). Arabian J. Chem. 10, 172–184.  Web of Science CSD CrossRef CAS Google Scholar
First citationZangrando, E., Islam, M. T., Islam, M. A. A. A., Sheikh, M. C., Tarafder, M. T. H., Miyatake, R., Zahan, R. & Hossain, M. A. (2015). Inorg. Chim. Acta, 427, 278–284.  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