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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

3,3-Bis(2-hy­dr­oxy­eth­yl)-1-(4-methyl­benzoyl)thio­urea: crystal structure, Hirshfeld surface analysis and computational study

CROSSMARK_Color_square_no_text.svg

aResearch Centre for Crystalline Materials, School of Science and Technology, Sunway University, 47500 Bandar Sunway, Selangor Darul Ehsan, Malaysia, and bDepartment of Physics, Bhavan's Sheth R. A. College of Science, Ahmedabad, Gujarat 380001, India
*Correspondence e-mail: edwardt@sunway.edu.my

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 9 September 2019; accepted 9 September 2019; online 12 September 2019)

In the title tri-substituted thio­urea derivative, C13H18N2O3S, the thione-S and carbonyl-O atoms lie, to a first approximation, to the same side of the mol­ecule [the S—C—N—C torsion angle is −49.3 (2)°]. The CN2S plane is almost planar (r.m.s. deviation = 0.018 Å) with the hy­droxy­ethyl groups lying to either side of this plane. One hy­droxy­ethyl group is orientated towards the thio­amide functionality enabling the formation of an intra­molecular N—H⋯O hydrogen bond leading to an S(7) loop. The dihedral angle [72.12 (9)°] between the planes through the CN2S atoms and the 4-tolyl ring indicates the mol­ecule is twisted. The experimental mol­ecular structure is close to the gas-phase, geometry-optimized structure calculated by DFT methods. In the mol­ecular packing, hydroxyl-O—H⋯O(hydrox­yl) and hydroxyl-O—H⋯S(thione) hydrogen bonds lead to the formation of a supra­molecular layer in the ab plane; no directional inter­actions are found between layers. The influence of the specified supra­molecular inter­actions is apparent in the calculated Hirshfeld surfaces and these are shown to be attractive in non-covalent inter­action plots; the inter­action energies point to the important stabilization provided by directional O—H⋯O hydrogen bonds.

1. Chemical context

The amine-H atoms in thio­urea, H2NC(=S)NH2, can be systematically replaced to generate up to tetra-functionalized mol­ecules, i.e. R1(R2)NC(=S)N(R3)R4 for R1–4 = alk­yl/aryl. The present study concerns a tri-substituted example, i.e. an N,N′-di(alk­yl/ar­yl)-N′-benzoyl­thio­urea derivative, notable for having a carbonyl group connected to the thio­urea framework. Thio­urea mol­ecules are of inter­est in themselves and as ligands for metal ions (Saeed et al., 2014[Saeed, A., Flörke, U. & Erben, M. F. (2014). J. Sulfur Chem. 35, 318-355.]). The free mol­ecules, including benzoyl derivatives, are well-known to exhibit various biological properties, for example, anti-bacterial, anti-fungal and anti-viral activities as well as cytotoxicity (Hallur et al., 2006[Hallur, G., Jimeno, A., Dalrymple, S., Zhu, T., Jung, M. K., Hidalgo, M., Isaacs, J. T., Sukumar, S., Hamel, E. & Khan, S. R. (2006). J. Med. Chem. 49, 2357-2360.]; Cunha et al., 2007[Cunha, S., Macedo, F. C., Costa, G. A. N., Rodrigues, M. T., Verde, R. B. V., de Souza Neta, L. C., Vencato, I., Lariucci, C. & Sá, F. P. (2007). Monatsh. Chem. 138, 511-516.]; Saeed et al., 2010[Saeed, S., Rashid, N., Jones, P. G., Ali, M. & Hussain, R. (2010). Eur. J. Med. Chem. 45, 1323-1331.]; Gunasekaran et al., 2017[Gunasekaran, N., Vadivel, V., Halcovitch, N. R. & Tiekink, E. R. T. (2017). Chem. Data Coll. 9-10, 263-276.]; Zhang et al., 2018[Zhang, Z.-J., Zeng, Y., Jiang, Z.-Y., Shu, B.-S., Sethuraman, V. & Zhong, G.-H. (2018). Pest Manag. Sci. 74, 1736-1746.]; Gunasekaran, Ng et al., 2012[Gunasekaran, N., Ng, S. W., Tiekink, E. R. T. & Karvembu, R. (2012). Polyhedron, 34, 41-45.]). The combination of hard (oxygen) and soft (sulfur) donor atoms along with nitro­gen suggests that benzoyl­thio­ureas can function as versatile ligands to metals. Indeed, a variety of coordination modes have been observed such as monodentate-S for the neutral ligand (Saeed et al., 2014[Saeed, A., Flörke, U. & Erben, M. F. (2014). J. Sulfur Chem. 35, 318-355.]; Gunasekaran, Ng et al., 2012[Gunasekaran, N., Ng, S. W., Tiekink, E. R. T. & Karvembu, R. (2012). Polyhedron, 34, 41-45.]). When deprotonated, a common mode of coordination is O-,S- chelation with considerable delocalization of π-electron density over the ensuing six-membered chelate ring (Saeed et al., 2014[Saeed, A., Flörke, U. & Erben, M. F. (2014). J. Sulfur Chem. 35, 318-355.]). While the motivations for preparing metal complexes of benzoyl­thio­ureas are varied, e.g. for anion recognition and as catalysts (Saeed et al., 2014[Saeed, A., Flörke, U. & Erben, M. F. (2014). J. Sulfur Chem. 35, 318-355.]; Zhang & Schreiner, 2009[Zhang, Z. & Schreiner, P. R. (2009). Chem. Soc. Rev. 38, 1187-1198.]; Nishikawa, 2018[Nishikawa, T. (2018). Tetrahedron Lett. 59, 216-223.]), there is continuing inter­est in exploring their biological potential as coordination of these ligands to metals generally enhances their biological efficacy, such as anti-cancer (Peng et al., 2016[Peng, B., Gao, Z., Li, X., Li, T., Chen, G., Zhou, M. & Zhang, J. (2016). J. Biol. Inorg. Chem. 21, 903-916.]; Barolli et al., 2017[Barolli, J. P., Maia, P. I. S., Colina-Vegas, L., Moreira, J., Plutin, A. M., Mocelo, R., Deflon, V. M., Cominetti, M. R., Camargo-Mathias, M. I. & Batista, A. A. (2017). Polyhedron, 126, 33-41.]; Jeyalakshmi et al., 2019[Jeyalakshmi, K., Haribabu, J., Balachandran, C., Narmatha, E., Bhuvanesh, N. S. P., Aoki, S., Awale, S. & Karvembu, R. (2019). New J. Chem. 43, 3188-3198.]), anti-microbial (Gemili et al., 2017[Gemili, M., Sari, H., Ulger, M., Sahin, E. & Nural, Y. (2017). Inorg. Chim. Acta, 463, 88-96.]; Binzet et al., 2018[Binzet, G., Gumus, I., Dogen, A., Flörke, U., Kulcu, N. & Arslan, H. (2018). J. Mol. Struct. 1161, 519-529.]; Saeed et al., 2018[Saeed, A., Larik, F. A., Jabeen, F., Mehfooz, H., Ghumro, S. A., El-Seedi, H. R., Ali, M., Channar, P. A. & Ashraf, H. (2018). Russ. J. Gen. Chem. 88, 541-550.]) and anti-mycobacterium tuberculosis (Plutín et al., 2016[Plutín, A. M., Alvarez, A., Mocelo, R., Ramos, R., Castellano, E. E., da Silva, M. M., Colina-Vegas, L., Pavan, F. R. & Batista, A. A. (2016). Inorg. Chem. Commun. 63, 74-80.]) activities. The present study was motivated by these applications and by previous structural studies (Gunasekaran et al., 2017[Gunasekaran, N., Vadivel, V., Halcovitch, N. R. & Tiekink, E. R. T. (2017). Chem. Data Coll. 9-10, 263-276.]; Selvakumaran & Karvembu, et al., 2011[Selvakumaran, N., Karvembu, R., Ng, S. W. & Tiekink, E. R. T. (2011). Acta Cryst. E67, o602.]; Selvakumaran, Ng et al., 2011[Selvakumaran, N., Ng, S. W., Tiekink, E. R. T. & Karvembu, R. (2011). Inorg. Chim. Acta, 376, 278-284.]) and the known catalytic applications of their cobalt complexes (Gunasekaran, Jerome et al., 2012[Gunasekaran, N., Jerome, P., Ng, S. W., Tiekink, E. R. T. & Karvembu, R. (2012). J. Molec. Catal. A: Chem. 353-354, 156-162.]). Herein, the synthesis, spectroscopic characterization and X-ray crystallographic investigation of the title compound, 4-MePhC(=O)N(H)C(=S)N(CH2CH2OH)2, (I)[link], are described, along with an analysis of the calculated Hirshfeld surfaces, non-covalent inter­action plots as well as a computational chemistry study.

[Scheme 1]

2. Structural commentary

The title compound, (I)[link], is illustrated in Fig. 1[link], and selected inter­atomic parameters are given in Table 1[link]. The structure features a tri-substituted thio­urea mol­ecule with one N atom bearing a benzoyl residue and the other, carrying two hy­droxy­ethyl groups. The thione-S and carbonyl-O atoms lie to the same side of the mol­ecule but are only approximately syn as the S1—C1—N2—C6 torsion angle is −49.3 (2)°; the O3—C6—N2—C1 torsion angle is −6.8 (3)°. The hy­droxy­ethyl groups lie to either side of the CN2S plane (r.m.s. deviation = 0.018 Å). The O1-hy­droxy­ethyl group is folded toward the thio­amide part of the mol­ecule, an orientation that allows for the formation of an intra­molecular N2—H⋯O1 hydrogen bond that closes an S(7) loop, Table 2[link]. Overall, the mol­ecule is twisted as seen in the dihedral angle of 72.12 (9)° between the CN2S atoms and the terminal aryl ring. The C1—N1 bond length is considerably shorter than the C1—N2 bond, which suggests some delocalization of π-electron density over the S1—C1—N1 atoms that does not extend over the C1—N1—C6 atoms, consistent with the large twist about the C1—N2 bond (see above). The bond angles subtended at the C1 and C6 atoms follow the expected trends in that those involving the formally doubly bonded atoms are wider, by approximately 10°, compared with the other angles, Table 1[link].

Table 1
Selected geometric parameters for (I)[link] determined experimentally (X-ray) and from theory

Parameter X-ray Theory
C1—S1 1.6744 (17) 1.671
C1—N1 1.335 (2) 1.368
C1—N2 1.396 (2) 1.404
C6—O3 1.214 (2) 1.220
C6—N2 1.382 (2) 1.396
S1—C1—N1 123.97 (13) 124.2
S1—C1—N2 121.67 (12) 122.2
N1—C1—N2 114.30 (14) 113.5
O3—C6—N2 122.09 (17) 123.3
O3—C6—C7 122.23 (16) 122.0
N2—C6—C7 115.64 (15) 114.7
S1—C1—N2—C6 −49.3 (2) −43.5
S1—C1—N1—C2 171.63 (12) 167.1
S1—C1—N1—C4 −7.3 (2) −7.6
O3—C6—N2—C1 −6.8 (3) −18.2
O3—C6—C7—C8 160.25 (17) 156.2
N1—C2—C3—O1 −70.1 (2) −69.2
N1—C4—C5—O2 57.5 (2) 69.0

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
N2—H2N⋯O1 0.87 (1) 1.91 (1) 2.728 (2) 157 (1)
O1—H1O⋯O2i 0.83 (2) 1.94 (2) 2.769 (2) 172 (2)
O2—H2O⋯S1ii 0.84 (2) 2.38 (2) 3.2049 (14) 171 (2)
C8—H8⋯O3iii 0.93 2.38 3.251 (2) 156
Symmetry codes: (i) [-x, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (ii) x-1, y, z; (iii) [-x+1, y-{\script{1\over 2}}, -z+{\script{1\over 2}}].
[Figure 1]
Figure 1
The mol­ecular structure of (I)[link] showing the atom-labelling scheme and displacement ellipsoids at the 50% probability level.

3. Gas-phase theoretical structure

Compound (I)[link] was subjected to gas-phase geometry optimization by long-range corrected wB97XD density functional with Grimme's D2 dispersion model (Chai & Head-Gordon, 2008[Chai, J. D. & Head-Gordon, M. (2008). Phys. Chem. Chem. Phys. 10, 6615-6620.]) coupled with Pople's 6-311+G(d,p) basis set (Petersson et al., 1988[Petersson, G. A., Bennett, A., Tensfeldt, T. G., Al-Laham, M. A., Shirley, W. A. & Mantzaris, J. (1988). J. Chem. Phys. 89, 2193-2218.]) as implemented in Gaussian16 (Frisch et al., 2016[Frisch, M. J., et al. (2016). Gaussian 16, Revision A. 03. Gaussian, Inc., Wallingford CT, USA.]) in order to compare the optimized mol­ecule with the experimental structure. The results of the optimization show that the local minimum structure in the gas-phase was located as confirmed through a frequency analysis with zero imaginary frequency. The superimposition of the experimental and theoretical structures (Macrae et al., 2006[Macrae, C. F., Edgington, P. R., McCabe, P., Pidcock, E., Shields, G. P., Taylor, R., Towler, M. & van de Streek, J. (2006). J. Appl. Cryst. 39, 453-457.]), Fig. 2[link], indicates that there are minor differences between the mol­ecules in either phase, with the r.m.s. deviation between them being 0.014 Å. Salient geometric data for the gas-phase structure are included in Table 1[link] and correlate very well with the experimental results. The major differences between the experimental and geometry-optimized structures relates to differences in the (i) O3—C6—N2—C1 torsion angles, which deviates further, by approximately 10°, from the anti-disposition in the optimized structure, and (ii) N1—C2—C3—O1 and N1—C4—C5—O2 torsion angles, which are disparate, by about 12°, in the experimental structure but are symmetric, i.e. ±69°, in the optimized structure.

[Figure 2]
Figure 2
Overlay diagram for experimental (green image) and geometry-optimized (red) mol­ecules for (I)[link]. The mol­ecules have been overlapped so the S=C—N—C=O fragments are coincident.

4. Supra­molecular features

In the crystal of (I)[link], the O1-hydroxyl group acts as a hydrogen-bond donor to the O2-hy­droxy group, which in turn functions as a donor to the S1-atom, Table 2[link]. The O—H⋯O hydrogen bonding is propagated by 21 symmetry to generate helical chains along the b-axis direction. The O—H⋯S hydrogen bonding serves to connect translationally related chains along the a-axis direction and these contacts are reinforced by phenyl-C—H⋯O(carbon­yl) inter­actions. In this way, a supra­molecular layer in the ab plane is formed, Fig. 3[link](a). Layers stack along the c-axis direction without directional inter­actions between them, Fig. 3[link](b).

[Figure 3]
Figure 3
Views of the mol­ecular packing in (I)[link]: (a) supra­molecular layer in the ab plane sustained by hy­droxy-O—H⋯O(hy­droxy), hy­droxy-O—H⋯S(thione) and aryl-C—H⋯O(carbon­yl) inter­actions, and (b) view of the unit-cell contents in projection down the a axis, highlighting the stacking of layers; one layer is represented in space-filling mode. The O—H⋯O, O—H⋯S and C–H⋯O inter­actions are shown as orange, blue and green dashed lines, respectively.

5. Hirshfeld surface analysis

The calculations of the Hirshfeld surfaces and the two-dimensional fingerprint plots (overall and delineated) for (I)[link] were performed using Crystal Explorer 17 (Turner et al., 2017[Turner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.]) and published protocols (Tan et al., 2019[Tan, S. L., Jotani, M. M. & Tiekink, E. R. T. (2019). Acta Cryst. E75, 308-318.]).

The Hirshfeld surface mapped over electrostatic potential in Fig. 4[link], shows different potentials surrounding the key functional groups. Thus, the donors and acceptors of conventional O—H⋯O and O—H⋯S hydrogen bonds and C—H⋯O contacts appear as blue and red regions, respectively, corresponding to positive and negative potential. The Hirshfeld surface mapped over dnorm in Fig. 5[link] also gives the usual indications of these inter­molecular inter­actions through the appearance of bright-red spots near participating atoms. In addition, short inter­atomic contacts between the hydroxyl-H atom, and carbonyl-C6 and hydroxyl-O2 atoms, and between the ethyl-C5 and hydroxyl-H1O atoms, Table 3[link], are either characterized as faint-red spots or merged within the bright-red spots corresponding to the conventional hydrogen bonds in Fig. 5[link].

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

The inter­atomic distances are calculated in Crystal Explorer 17 (Turner et al., 2017[Turner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.]) whereby the X—H bond lengths are adjusted to their neutron values.

Contact Distance Symmetry operation
H1O⋯H2O 2.26 x, −[{1\over 2}] + y, −[{1\over 2}] − z
C3⋯O3 3.112 (2) x, −[{1\over 2}] + y, [{1\over 2}] − z
H2B⋯O3 2.58 x, −[{1\over 2}] + y, [{1\over 2}] − z
H3B⋯O3 2.69 x, −[{1\over 2}] + y, [{1\over 2}] − z
C5⋯H1O 2.73 x, −[{1\over 2}] + y, [{1\over 2}] − z
H13A⋯O1 2.67 1 − x, 1 − y, −z
C6⋯O2 3.177 (2) 1 + x, y, z
C8⋯H2B 2.78 1 + x, y, z
[Figure 4]
Figure 4
A view of the Hirshfeld surface mapped over the calculated electrostatic potential for (I)[link]. The red and blue regions represent negative and positive electrostatic potentials, respectively. The potentials were calculated using the STO-3G basis set at Hartree–Fock level of theory over a range of ±0.18 atomic units.
[Figure 5]
Figure 5
Two views of the Hirshfeld surface mapped over dnorm for (I)[link] in the range −0.132 to +1.682 arbitrary units.

The inter­molecular contacts in the crystal of (I)[link] were further analysed using an enrichment ratio (ER) descriptor, which is derived from the analysis of the Hirshfeld surface (Jelsch et al., 2014[Jelsch, C., Ejsmont, K. & Huder, L. (2014). IUCrJ, 1, 119-128.]). The ER relates the propensity of pair of chemical species to form a specific inter­action in a crystal. The enrichment ratio, ER(X, Y), for a pair of elements (X, Y) is defined as the ratio between proportion of actual contacts in the crystal to the theoretical proportion of random contacts. This ratio is greater than unity for a pair of elements having a high likelihood to form contacts in a crystal, while it is less than one for a pair which tends to avoid contacts with each other. A listing of ER values for (I)[link] is given in Table 4[link]. The enrichment ratios greater than unity for the atom pairs (O, H) and (S, H), Table 4[link], are consistent with the high propensity for the formation of the O—H⋯O and O—H⋯S hydrogen bonds in the crystal. It is also evident that the value greater than unity for (C, H) arises from the C⋯H/H⋯C contacts.

Table 4
Enrichment ratios for (I)

Parameter Ratio
H⋯H 0.92
C⋯H 1.21
O⋯H 1.21
S⋯H 1.33
C⋯O 0.54

The overall fingerprint plots for (I)[link] and those delineated into H⋯H, O⋯H/H⋯O, C⋯H/H⋯C and S⋯H/H⋯S contacts are illustrated in Fig. 6[link](a)–(e), respectively. A summary of the percentage contributions from the various contacts in the crystal are given in Table 5[link]. The contribution from H⋯H contacts are reflected in the middle of the scattered point and cover the greatest area in the plot, and make the most significant contribution (52.5%) to the total Hirshfeld surface, Fig. 6[link](b) and has an ER value of 0.92, i.e. close to unity. The contribution from O⋯H/H⋯O contacts is viewed as long spikes at de + di ∼1.8 Å, with points scattered around different regions in the delineated fingerprint plot, Fig. 6[link](c). In the fingerprint delineated into C⋯H/H⋯C contacts in Fig. 6[link](d), a pair of small tips at de + di < 2.8 Å is the result of short inter­atomic contacts, Table 3[link], including an inter­layer contact (H13A⋯O1). The percentage contribution from S⋯H/H⋯S contacts (13.1%) reflect the presence of O—H⋯S hydrogen bonds and are apparent through the appearance of asymmetric spikes at de + di ∼2.1 Å in Fig. 6[link](e).

Table 5
Percentage contributions of inter­atomic contacts to the Hirshfeld surface for (I)

Contact Percentage contribution
H⋯H 52.5
C⋯H/H⋯C 16.2
O⋯H/H⋯O 15.0
S⋯H/H⋯S 13.1
N⋯H/H⋯N 1.5
C⋯C 0.3
C⋯O/O⋯C 0.8
N⋯O/O⋯N 0.1
O⋯O 0.3
C⋯N/N⋯C 0.2
[Figure 6]
Figure 6
(a) A comparison of the full two-dimensional fingerprint plot for (I)[link] and those delineated into (b) H⋯H, (c) O⋯H/H⋯O, (d) C⋯H/H⋯C and (e) S⋯H/H⋯S contacts.

6. Computational chemistry

The inter­molecular O—H⋯O, O—H⋯S and C—H⋯O inter­actions occurring between the respective pairs of mol­ecules were subjected to energy calculations by DFT-wB97XD/aug-cc-pVTZ (Woon & Dunning, 1993[Woon, D. E. & Dunning, T. H. Jr (1993). J. Chem. Phys. 98, 1358-1371.]) for the evaluation of the strength of these inter­actions. With reference to the BSSE corrected inter­action energies (EBSSEint) listed in Table 6[link], the O—H⋯O hydrogen bond has the greatest inter­action energy, followed by C—H⋯O and O—H⋯S. Unexpectedly, the C—H⋯O inter­action has an energy approximately 3–4 kcal mol−1 more stable than the O—H⋯S inter­action despite phenyl-C—H being a weak hydrogen-bond donor and thione-S a weak acceptor, and that such inter­actions are known to be dispersive in nature (Bhattacharyya et al., 2013[Bhattacharyya, S., Bhattacherjee, A., Shirhatti, P. R. & Wategaonkar, S. (2013). J. Phys. Chem. A, 117, 8238-8250.]). The donor–acceptor inter­actions were also evaluated by a natural bond orbital (NBO) population analysis (Reed et al., 1988[Reed, A. E., Curtiss, L. A. & Weinhold, F. (1988). Chem. Rev. 88, 899-926.]), which revealed that the net NBO charge for H8⋯O3 is 0.8 compared to 0.6 for H2O⋯S1, thereby confirming the relative strength of these inter­actions.

Table 6
Summary of inter­action energies (kcal mol−1) calculated for several directional contacts in (I)

Contact Etot
O1—H1O⋯O2 −14.52
O2—H2O⋯S1 −6.27
C8—H8⋯O3 −9.65

To complement the results of the calculations on the inter­action energies, the dimeric structures were subjected to further analysis by NCIPLOT (Johnson et al., 2010[Johnson, E. R., Keinan, S., Mori-Sánchez, P., Contreras-García, J., Cohen, A. J. & Yang, W. (2010). J. Am. Chem. Soc. 132, 6498-6506.]). The analysis provides a convenient visualization index on the strength of any existing non-covalent inter­actions through a red–blue–green colour scheme on the isosurface, i.e. red is indicative of a strong repulsive inter­action, blue is indicative of strong attractive inter­action while green is indicative of a weak inter­action (Contreras-García et al., 2011[Contreras-García, J., Johnson, E. R., Keinan, S., Chaudret, R., Piquemal, J. P., Beratan, D. N. & Yang, W. (2011). J. Chem. Theory Comput. 7, 625-632.]). The results, illus­trated in Fig. 7[link], reveal that the O—H⋯O inter­action is clearly strong and attractive, while both O—H⋯S and C—H⋯O are considered weak inter­actions.

[Figure 7]
Figure 7
NCI plots for the dimeric aggregates in (I)[link] sustained by (a) O—H⋯O, (b) O—H⋯S and (c) C—H⋯O inter­actions (highlighted in boxes), and (d) plot of RDG versus sign(λ2)ρ(r). The gradient cut-off is set at 0.4 and the colour scale is −0.03 < ρ < 0.03 atomic units.

As the mol­ecular packing is governed directionally by hydrogen bonding between mol­ecules, the energy frameworks were simulated (Turner et al., 2017[Turner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.]) in order to compare the topology of these inter­molecular inter­actions. A detailed analysis of the energy frameworks shown in Fig. 8[link] reveals the crystal of (I)[link] is mainly stabilized by electrostatic and dispersive forces. The total electrostatic energy (Eelectrostatic) of all pairwise inter­actions sums to −36.11 kcal mol−1, while the total dispersion energy term (Edispersion) computes to −43.83 kcal mol−1.

[Figure 8]
Figure 8
The energy framework diagrams for (I)[link] showing (a) Eelectrostatic (red cylinders), (b) Edispersion (green cylinders) and (c) Etotal (blue cylinders), viewed along the a axis. The frameworks were adjusted to the same scale factor of 50 with a cut-off value of 2.39 kcal mol−1 within 2 × 2 × 2 unit cells. The corresponding cylinder radii are proportional to the relative magnitude of the energies.

7. Database survey

The crystal structure of the parent compound, PhC(=O)N(H)C(=S)N(CH2CH2OH)2, (II), has been reported twice (Koch et al., 1995[Koch, K. R., Sacht, C. & Bourne, S. (1995). Inorg. Chim. Acta, 232, 109-115.]; Cornejo et al., 2005[Cornejo, J. A., Ayala, K., Richter, R., Böhlig, H., Hennig, L. & Beyer, L. (2005). Z. Anorg. Allg. Chem. 631, 3040-3045.]; refcodes ZAJWAI and ZAJWAI01, respectively). The conformation of this mol­ecule and that of (I)[link] are very similar and the geometric parameters describing chemically equivalent parameters are generally within experimental errors. The most important conformational difference is seen in the pair of N1—C2—C3—O1 [73.7 (2)°] and N1—C4—C5—O1 [–53.9 (2)°] torsion angles, which span a range of approximately 20° in (II) cf. approximately 12° in (I)[link]. The mol­ecular packing in (II) also features O—H⋯O hydrogen and O—H⋯S hydrogen bonding, as for (I)[link], leading to a supra­molecular layer; the intra­molecular amine-N—H⋯·O(hy­droxy) hydrogen bond persists. However, in the case of (II), there are directional inter­actions between layers, i.e. of the type phenyl-C—H⋯π(phen­yl), to sustain a three-dimensional architecture. The other closely related structure is that of 4-MePhC(=O)N(H)C(=S)N(Me)CH2CH2OH) (Jamaludin et al., 2016[Jamaludin, N. S., Halim, S. N. A. & Tiekink, E. R. T. (2016). IUCrData, 1, x152457.]; refcode GADBOF). Here, the intra­molecular amine-N—H⋯O(hy­droxy) hydrogen bond is also found and the most prominent feature of the mol­ecular packing is the formation of supra­molecular helical chains mediated by hy­droxy-O—H⋯O(carbon­yl) hydrogen bonds.

8. Synthesis and crystallization

All chemicals and solvents were used as purchased without purification. The reactions were carried out under ambient conditions. The melting point was measured using a Hanon MP-450 melting point apparatus. The CHN elemental analysis was performed on a LECO TruSpec Micro analyser under helium atmosphere with glycine being used as the standard. The IR spectrum was measured on a Bruker Vertex 70v FT–IR spectrophotometer from 4000 to 400 cm−1. The 1H and 13C{1H} spectra were recorded in DMSO-d6 solutions on a Bruker Ascend 400 MHz NMR spectrometer with chemical shifts relative to tetra­methyl­silane (TMS). The optical absorption spectra were measured on 10 and 100 µM ethanol:aceto­nitrile (1:1) solutions in the range 190–1100 nm on a double-beam Shimadzu UV 3600 Plus UV–vis spectrophotometer. The thermogravimetric analysis (TGA) was performed on a Perkin Elmer STA 6000 Simultaneous Thermogravimetric Analyzer in the range of 35–900°C under a nitro­gen atmosphere at a flow rate of 10°C min−1. The experimental powder X-ray diffraction pattern was measured on a Rigaku MiniFlex diffractometer with Cu Kα1 radiation (λ = 1.54056 Å) in the 2θ range of 5–70° and a step size of 0.02°. The experimental PXRD patterns were compared to the simulated PXRD patterns calculated from the CIF using the Rigaku PDXL structure analysis software package. The patterns matched indicating that the reported crystal structure is representative of the bulk material.

Synthesis of (I)[link]: An excess of thionyl chloride (Merck) was mixed with 4-methyl­benzoic acid (Merck, 1 mmol) and the resulting solution was refluxed until a pale-yellow solution was obtained. The excess thionyl chloride was removed on a water bath, leaving only 4-methyl­benzoyl chloride, which is a yellow, viscous liquid. Ammonium thio­cyanate (Fisher, 1 mmol) was added into an acetone (30 ml) solution of 4-methyl­benzoyl chloride (1 mmol). The solution turned yellow after stirring for 2 h. The white precipitate (ammonium chloride) was isolated upon filtration and to the yellow filtrate, bis­(hy­droxy­eth­yl)amine (Acros, 1 mmol) was carefully added followed by stirring for 1 h. Upon the addition of di­chloro­methane (50 ml), a yellow precipitate was obtained, which was collected by filtration. Recrystallization from its hot acetone solution yielded colourless blocks after slow evaporation. White solid, yield 56%, m.p. 400.3–402.1 K. Elemental analysis: C13H18N2O3S, found (calculated): C 55.59 (55.30), H 6.57 (6.43), N 9.79 (9.92). IR (ATR; cm−1): 3312 (br, νOH), 3158 (br, νNH), 3061 (w, νCHaro), 2955–2881 (w, νCH), 1686 (s, νC=O), 1539 (s, νC=C), 1250 (s, νC—N), 1054 (s, νC=S), 747 (s, δCH). UV (ethanol:aceto­nitrile; 5ml:5ml): λmax nm (assignment; log ) 354.4 (nπ*; 4.34), 294.0 (nπ*; 4.98), 246.4 (ππ*; 5.17), 202.6 (σπ*; 5.17). 1H NMR (400 MHz, DMSO-d6; see Fig. 1[link] for the numbering scheme): δ 10.78 (1H, br, s, NH), 7.76 (2H, d, 2-phenyl, 3JHH = 7.72 Hz), 7.31 (2H, d, 3-phenyl, 3JHH = 7.6 Hz), 5.66 (1H, br, s, OH), 4.87 (1H, t, OH, 3JOH–H = 5.00 Hz), 3.98 (2H, overlapping t, CH2–C2, 3JHH = 6.24 Hz, 3JHH = 6.08 Hz), 3.76 (2H, m, CH2–C3), 3.70 (4H, m, CH2–C4, C5), 2.37 (3H, s, CH3). 13C{1H} NMR (100 MHz, DMSO-d6): δ 180.63 (C1), 163.78 (C6), 141.88 (C7), 130.13 (C10), 128.47 (C9, C11), 127.28 (C8, C12), 58.58 (C5), 56.95 (C3), 54.42 (C4), 54.29 (C2), 20.42 (C13).

The pyrolytic processes for (I)[link] was resolved into four main stages. The first stage involves the liberation of H2O between 135 and 165°C, which corresponds to approximate 6% of the weight for (I)[link]. The second stage between 160 and 240°C is attributed to the loss of a 4-methyl­benzaldehyde fragment, corresponding to 45% weight loss. Subsequently, the remaining fragments undergo further pyrolysis to result in the liberation of ethanol (31% weight) and ammonia (17–18%) in the range 230 to 300°C and 300°C onward, respectively. Compound (I)[link] decomposed at temperatures beyond 700°C.

9. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 7[link]. Carbon-bound H atoms were placed in calculated positions (C—H = 0.93–0.97 Å) and were included in the refinement in the riding-model approximation, with Uiso(H) set to 1.2–1.5Ueq(C). The O- and N-bound H atoms were located from a difference map and refined with O—H and N—H = 0.84±0.01 and 0.88±0.01 Å, respectively, and with Uiso(H) = 1.5Ueq(O) and 1.2Ueq(N).

Table 7
Experimental details

Crystal data
Chemical formula C13H18N2O3S
Mr 282.35
Crystal system, space group Monoclinic, P21/c
Temperature (K) 293
a, b, c (Å) 7.4051 (10), 10.6213 (15), 18.569 (3)
β (°) 94.117 (2)
V3) 1456.7 (4)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.23
Crystal size (mm) 0.12 × 0.09 × 0.08
 
Data collection
Diffractometer Bruker SMART APEX
Absorption correction Multi-scan (SADABS; Sheldrick, 1996[Sheldrick, G. M. (1996). SADABS. University of Göttingen, Germany.])
Tmin, Tmax 0.655, 0.746
No. of measured, independent and observed [I > 2σ(I)] reflections 18125, 3339, 2263
Rint 0.051
(sin θ/λ)max−1) 0.650
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.039, 0.098, 1.04
No. of reflections 3339
No. of parameters 182
No. of restraints 3
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.17, −0.20
Computer programs: SMART and SAINT (Bruker, 2008[Bruker (2008). SMART and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2014/7 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), 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: SMART (Bruker, 2008); cell refinement: SMART (Bruker, 2008); data reduction: SAINT (Bruker, 2008); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008); program(s) used to refine structure: SHELXL2014/7 (Sheldrick, 2015); molecular graphics: ORTEP-3 for Windows (Farrugia, 2012), DIAMOND (Brandenburg, 2006); software used to prepare material for publication: publCIF (Westrip, 2010).

3,3-Bis(2-hydroxyethyl)-1-(4-methylbenzoyl)thiourea top
Crystal data top
C13H18N2O3SF(000) = 600
Mr = 282.35Dx = 1.287 Mg m3
Monoclinic, P21/cMo Kα radiation, λ = 0.71073 Å
a = 7.4051 (10) ÅCell parameters from 2509 reflections
b = 10.6213 (15) Åθ = 2.2–22.9°
c = 18.569 (3) ŵ = 0.23 mm1
β = 94.117 (2)°T = 293 K
V = 1456.7 (4) Å3Prism, colourless
Z = 40.12 × 0.09 × 0.08 mm
Data collection top
Bruker SMART APEX
diffractometer
3339 independent reflections
Radiation source: fine-focus sealed tube2263 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.051
φ and ω scansθmax = 27.5°, θmin = 2.2°
Absorption correction: multi-scan
(SADABS; Sheldrick, 1996)
h = 99
Tmin = 0.655, Tmax = 0.746k = 1313
18125 measured reflectionsl = 2424
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)] = 0.039Hydrogen site location: mixed
wR(F2) = 0.098H atoms treated by a mixture of independent and constrained refinement
S = 1.03 w = 1/[σ2(Fo2) + (0.0374P)2 + 0.2262P]
where P = (Fo2 + 2Fc2)/3
3339 reflections(Δ/σ)max < 0.001
182 parametersΔρmax = 0.17 e Å3
3 restraintsΔρmin = 0.20 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
S10.32019 (6)0.58997 (5)0.40517 (3)0.05150 (16)
O10.20850 (18)0.31841 (13)0.20234 (8)0.0600 (4)
H1O0.239 (3)0.2473 (13)0.1888 (13)0.090*
O20.27874 (16)0.58125 (13)0.35097 (7)0.0534 (4)
H2O0.3778 (19)0.585 (2)0.3699 (11)0.080*
O30.37968 (18)0.75187 (12)0.26514 (8)0.0602 (4)
N10.06795 (17)0.47500 (13)0.31953 (7)0.0382 (3)
N20.31599 (19)0.54248 (13)0.26306 (8)0.0402 (3)
H2N0.304 (2)0.4767 (12)0.2356 (8)0.048*
C10.2270 (2)0.53452 (15)0.32668 (9)0.0380 (4)
C20.0239 (2)0.44158 (17)0.24885 (10)0.0448 (4)
H2A0.00530.50450.21370.054*
H2B0.15370.44380.25280.054*
C30.0273 (2)0.31392 (18)0.22177 (11)0.0524 (5)
H3A0.01550.25130.25920.063*
H3B0.05260.29070.18020.063*
C40.0291 (2)0.44025 (17)0.38245 (10)0.0456 (4)
H4A0.05830.41790.42180.055*
H4B0.10240.36640.37070.055*
C50.1497 (2)0.54346 (19)0.40748 (10)0.0497 (5)
H5A0.21250.51380.44830.060*
H5B0.07620.61530.42320.060*
C60.3952 (2)0.64952 (16)0.23710 (9)0.0405 (4)
C70.4941 (2)0.63211 (16)0.17111 (9)0.0393 (4)
C80.5608 (2)0.51755 (17)0.14953 (9)0.0449 (4)
H80.54190.44550.17640.054*
C90.6557 (3)0.5097 (2)0.08809 (10)0.0562 (5)
H90.70110.43220.07470.067*
C100.6843 (3)0.6145 (2)0.04626 (10)0.0584 (5)
C110.6157 (3)0.7282 (2)0.06798 (11)0.0631 (6)
H110.63200.79990.04040.076*
C120.5241 (3)0.73742 (18)0.12945 (11)0.0557 (5)
H120.48150.81540.14340.067*
C130.7870 (4)0.6055 (3)0.02119 (12)0.0919 (9)
H13A0.72240.65060.05980.138*
H13B0.79830.51870.03460.138*
H13C0.90520.64160.01200.138*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
S10.0407 (3)0.0703 (4)0.0435 (3)0.0051 (2)0.00312 (19)0.0119 (2)
O10.0472 (8)0.0542 (9)0.0806 (10)0.0075 (7)0.0181 (7)0.0248 (7)
O20.0382 (7)0.0691 (9)0.0540 (8)0.0084 (6)0.0118 (6)0.0126 (7)
O30.0621 (9)0.0393 (8)0.0818 (10)0.0049 (6)0.0227 (7)0.0135 (7)
N10.0332 (7)0.0397 (8)0.0423 (8)0.0014 (6)0.0059 (6)0.0025 (6)
N20.0438 (8)0.0378 (8)0.0399 (8)0.0097 (7)0.0093 (6)0.0063 (6)
C10.0357 (9)0.0356 (9)0.0432 (10)0.0024 (7)0.0054 (7)0.0012 (7)
C20.0370 (9)0.0481 (11)0.0487 (10)0.0027 (8)0.0003 (8)0.0064 (8)
C30.0458 (11)0.0512 (12)0.0613 (12)0.0102 (9)0.0114 (9)0.0150 (9)
C40.0400 (10)0.0482 (11)0.0491 (11)0.0029 (8)0.0068 (8)0.0090 (8)
C50.0430 (10)0.0635 (12)0.0435 (10)0.0005 (9)0.0097 (8)0.0033 (9)
C60.0346 (9)0.0376 (10)0.0491 (10)0.0020 (7)0.0019 (8)0.0023 (8)
C70.0356 (9)0.0400 (10)0.0419 (9)0.0061 (7)0.0003 (7)0.0023 (7)
C80.0481 (10)0.0430 (10)0.0443 (10)0.0006 (8)0.0073 (8)0.0073 (8)
C90.0598 (13)0.0569 (13)0.0530 (12)0.0013 (10)0.0121 (10)0.0038 (10)
C100.0558 (12)0.0773 (16)0.0425 (11)0.0167 (11)0.0067 (9)0.0046 (10)
C110.0703 (14)0.0617 (14)0.0577 (13)0.0171 (11)0.0080 (11)0.0206 (11)
C120.0600 (12)0.0422 (11)0.0653 (13)0.0075 (9)0.0076 (10)0.0085 (9)
C130.0952 (19)0.126 (2)0.0581 (14)0.0220 (17)0.0303 (14)0.0050 (14)
Geometric parameters (Å, º) top
S1—C11.6744 (17)C4—H4B0.9700
O1—C31.415 (2)C5—H5A0.9700
O1—H1O0.832 (9)C5—H5B0.9700
O2—C51.425 (2)C6—C71.483 (2)
O2—H2O0.836 (9)C7—C81.383 (2)
O3—C61.214 (2)C7—C121.387 (2)
N1—C11.335 (2)C8—C91.384 (2)
N1—C41.462 (2)C8—H80.9300
N1—C21.477 (2)C9—C101.382 (3)
N2—C61.382 (2)C9—H90.9300
N2—C11.396 (2)C10—C111.382 (3)
N2—H2N0.866 (9)C10—C131.514 (3)
C2—C31.504 (2)C11—C121.373 (3)
C2—H2A0.9700C11—H110.9300
C2—H2B0.9700C12—H120.9300
C3—H3A0.9700C13—H13A0.9600
C3—H3B0.9700C13—H13B0.9600
C4—C51.508 (3)C13—H13C0.9600
C4—H4A0.9700
C3—O1—H1O109.1 (18)C4—C5—H5A109.4
C5—O2—H2O105.3 (16)O2—C5—H5B109.4
C1—N1—C4121.39 (14)C4—C5—H5B109.4
C1—N1—C2123.27 (14)H5A—C5—H5B108.0
C4—N1—C2115.33 (13)O3—C6—N2122.09 (17)
C6—N2—C1125.67 (14)O3—C6—C7122.23 (16)
C6—N2—H2N119.0 (12)N2—C6—C7115.64 (15)
C1—N2—H2N114.6 (12)C8—C7—C12118.21 (17)
N1—C1—N2114.30 (14)C8—C7—C6123.83 (16)
N1—C1—S1123.97 (13)C12—C7—C6117.95 (16)
N2—C1—S1121.67 (12)C7—C8—C9120.32 (17)
N1—C2—C3113.76 (15)C7—C8—H8119.8
N1—C2—H2A108.8C9—C8—H8119.8
C3—C2—H2A108.8C10—C9—C8121.45 (19)
N1—C2—H2B108.8C10—C9—H9119.3
C3—C2—H2B108.8C8—C9—H9119.3
H2A—C2—H2B107.7C11—C10—C9117.76 (18)
O1—C3—C2108.71 (15)C11—C10—C13120.9 (2)
O1—C3—H3A109.9C9—C10—C13121.4 (2)
C2—C3—H3A109.9C12—C11—C10121.24 (19)
O1—C3—H3B109.9C12—C11—H11119.4
C2—C3—H3B109.9C10—C11—H11119.4
H3A—C3—H3B108.3C11—C12—C7121.01 (19)
N1—C4—C5113.48 (14)C11—C12—H12119.5
N1—C4—H4A108.9C7—C12—H12119.5
C5—C4—H4A108.9C10—C13—H13A109.5
N1—C4—H4B108.9C10—C13—H13B109.5
C5—C4—H4B108.9H13A—C13—H13B109.5
H4A—C4—H4B107.7C10—C13—H13C109.5
O2—C5—C4111.00 (15)H13A—C13—H13C109.5
O2—C5—H5A109.4H13B—C13—H13C109.5
C4—N1—C1—N2169.93 (14)O3—C6—C7—C8160.25 (17)
C2—N1—C1—N211.1 (2)N2—C6—C7—C822.1 (2)
C4—N1—C1—S17.3 (2)O3—C6—C7—C1218.5 (3)
C2—N1—C1—S1171.63 (12)N2—C6—C7—C12159.18 (16)
C6—N2—C1—N1133.37 (17)C12—C7—C8—C90.3 (3)
C6—N2—C1—S149.3 (2)C6—C7—C8—C9178.41 (16)
C1—N1—C2—C389.8 (2)C7—C8—C9—C100.9 (3)
C4—N1—C2—C391.18 (18)C8—C9—C10—C110.4 (3)
N1—C2—C3—O170.1 (2)C8—C9—C10—C13179.6 (2)
C1—N1—C4—C586.72 (19)C9—C10—C11—C120.7 (3)
C2—N1—C4—C592.28 (18)C13—C10—C11—C12179.3 (2)
N1—C4—C5—O257.5 (2)C10—C11—C12—C71.3 (3)
C1—N2—C6—O36.8 (3)C8—C7—C12—C110.7 (3)
C1—N2—C6—C7175.48 (15)C6—C7—C12—C11179.56 (18)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
N2—H2N···O10.87 (1)1.91 (1)2.728 (2)157 (1)
O1—H1O···O2i0.83 (2)1.94 (2)2.769 (2)172 (2)
O2—H2O···S1ii0.84 (2)2.38 (2)3.2049 (14)171 (2)
C8—H8···O3iii0.932.383.251 (2)156
Symmetry codes: (i) x, y1/2, z+1/2; (ii) x1, y, z; (iii) x+1, y1/2, z+1/2.
Selected geometric parameters for (I) determined experimentally (X-ray) and from theory top
ParameterX-rayTheory
C1—S11.6744 (17)1.671
C1—N11.335 (2)1.368
C1—N21.396 (2)1.404
C6—O31.214 (2)1.220
C6—N21.382 (2)1.396
S1—C1—N1123.97 (13)124.2
S1—C1—N2121.67 (12)122.2
N1—C1—N2114.30 (14)113.5
O3—C6—N2122.09 (17)123.3
O3—C6—C7122.23 (16)122.0
N2—C6—C7115.64 (15)114.7
S1—C1—N2—C6-49.3 (2)-43.5
S1—C1—N1—C2171.63 (12)167.1
S1—C1—N1—C4-7.3 (2)-7.6
O3—C6—N2—C1-6.8 (3)-18.2
O3—C6—C7—C8160.25 (17)156.2
N1—C2—C3—O1-70.1 (2)-69.2
N1—C4—C5—O257.5 (2)69.0
Summary of short interatomic contacts (Å) in (I) top
The interatomic distances are calculated in Crystal Explorer 17 (Turner et al., 2017) whereby the X—H bond lengths are adjusted to their neutron values.
ContactDistanceSymmetry operation
H1O···H2O2.26-x, -1/2 + y, -1/2 - z
C3···O33.112 (2)-x, -1/2 + y, 1/2 - z
H2B···O32.58-x, -1/2 + y, 1/2 - z
H3B···O32.69-x, -1/2 + y, 1/2 - z
C5···H1O2.73-x, -1/2 + y, 1/2 - z
H13A···O12.671 - x, 1 - y, -z
C6···O23.177 (2)1 + x, y, z
C8···H2B2.781 + x, y, z
Enrichment ratios for (I) top
ParameterRatio
H···H0.92
C···H1.21
O···H1.21
S···H1.33
C···O0.54
Percentage contributions of interatomic contacts to the Hirshfeld surface for (I) top
ContactPercentage contribution
H···H52.5
C···H/H···C16.2
O···H/H···O15.0
S···H/H···S13.1
N···H/H···N1.5
C···C0.3
C···O/O···C0.8
N···O/O···N0.1
O···O0.3
C···N/N···C0.2
Summary of interaction energies (kcal mol-1) calculated for several directional contacts in (I) top
ContactEtot
O1—H1O···O2-14.52
O2—H2O···S1-6.27
C8—H8···O3-9.65
 

Funding information

Crystallographic research at Sunway University is supported by Sunway University Sdn Bhd (grant No. STR-RCTR-RCCM-001-2019).

References

First citationBarolli, J. P., Maia, P. I. S., Colina-Vegas, L., Moreira, J., Plutin, A. M., Mocelo, R., Deflon, V. M., Cominetti, M. R., Camargo-Mathias, M. I. & Batista, A. A. (2017). Polyhedron, 126, 33–41.  Web of Science CSD CrossRef CAS Google Scholar
First citationBhattacharyya, S., Bhattacherjee, A., Shirhatti, P. R. & Wategaonkar, S. (2013). J. Phys. Chem. A, 117, 8238–8250.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBinzet, G., Gumus, I., Dogen, A., Flörke, U., Kulcu, N. & Arslan, H. (2018). J. Mol. Struct. 1161, 519–529.  Web of Science CSD CrossRef CAS Google Scholar
First citationBrandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBruker (2008). SMART and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationChai, J. D. & Head-Gordon, M. (2008). Phys. Chem. Chem. Phys. 10, 6615–6620.  Web of Science CrossRef PubMed CAS Google Scholar
First citationContreras-García, J., Johnson, E. R., Keinan, S., Chaudret, R., Piquemal, J. P., Beratan, D. N. & Yang, W. (2011). J. Chem. Theory Comput. 7, 625–632.  Web of Science PubMed Google Scholar
First citationCornejo, J. A., Ayala, K., Richter, R., Böhlig, H., Hennig, L. & Beyer, L. (2005). Z. Anorg. Allg. Chem. 631, 3040–3045.  Web of Science CSD CrossRef CAS Google Scholar
First citationCunha, S., Macedo, F. C., Costa, G. A. N., Rodrigues, M. T., Verde, R. B. V., de Souza Neta, L. C., Vencato, I., Lariucci, C. & Sá, F. P. (2007). Monatsh. Chem. 138, 511–516.  Web of Science CSD 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 citationFrisch, M. J., et al. (2016). Gaussian 16, Revision A. 03. Gaussian, Inc., Wallingford CT, USA.  Google Scholar
First citationGemili, M., Sari, H., Ulger, M., Sahin, E. & Nural, Y. (2017). Inorg. Chim. Acta, 463, 88–96.  Web of Science CSD CrossRef CAS Google Scholar
First citationGunasekaran, N., Jerome, P., Ng, S. W., Tiekink, E. R. T. & Karvembu, R. (2012). J. Molec. Catal. A: Chem. 353–354, 156–162.  Google Scholar
First citationGunasekaran, N., Ng, S. W., Tiekink, E. R. T. & Karvembu, R. (2012). Polyhedron, 34, 41–45.  Web of Science CSD CrossRef CAS Google Scholar
First citationGunasekaran, N., Vadivel, V., Halcovitch, N. R. & Tiekink, E. R. T. (2017). Chem. Data Coll. 9-10, 263–276.  Google Scholar
First citationHallur, G., Jimeno, A., Dalrymple, S., Zhu, T., Jung, M. K., Hidalgo, M., Isaacs, J. T., Sukumar, S., Hamel, E. & Khan, S. R. (2006). J. Med. Chem. 49, 2357–2360.  Web of Science CrossRef PubMed CAS Google Scholar
First citationJamaludin, N. S., Halim, S. N. A. & Tiekink, E. R. T. (2016). IUCrData, 1, x152457.  Google Scholar
First citationJelsch, C., Ejsmont, K. & Huder, L. (2014). IUCrJ, 1, 119–128.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationJeyalakshmi, K., Haribabu, J., Balachandran, C., Narmatha, E., Bhuvanesh, N. S. P., Aoki, S., Awale, S. & Karvembu, R. (2019). New J. Chem. 43, 3188–3198.  Web of Science CSD CrossRef CAS Google Scholar
First citationJohnson, E. R., Keinan, S., Mori-Sánchez, P., Contreras-García, J., Cohen, A. J. & Yang, W. (2010). J. Am. Chem. Soc. 132, 6498–6506.  Web of Science CrossRef CAS PubMed Google Scholar
First citationKoch, K. R., Sacht, C. & Bourne, S. (1995). Inorg. Chim. Acta, 232, 109–115.  CSD CrossRef CAS Web of Science Google Scholar
First citationMacrae, C. F., Edgington, P. R., McCabe, P., Pidcock, E., Shields, G. P., Taylor, R., Towler, M. & van de Streek, J. (2006). J. Appl. Cryst. 39, 453–457.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNishikawa, T. (2018). Tetrahedron Lett. 59, 216–223.  Web of Science CrossRef CAS Google Scholar
First citationPeng, B., Gao, Z., Li, X., Li, T., Chen, G., Zhou, M. & Zhang, J. (2016). J. Biol. Inorg. Chem. 21, 903–916.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPetersson, G. A., Bennett, A., Tensfeldt, T. G., Al–Laham, M. A., Shirley, W. A. & Mantzaris, J. (1988). J. Chem. Phys. 89, 2193–2218.  CrossRef CAS Web of Science Google Scholar
First citationPlutín, A. M., Alvarez, A., Mocelo, R., Ramos, R., Castellano, E. E., da Silva, M. M., Colina-Vegas, L., Pavan, F. R. & Batista, A. A. (2016). Inorg. Chem. Commun. 63, 74–80.  Google Scholar
First citationReed, A. E., Curtiss, L. A. & Weinhold, F. (1988). Chem. Rev. 88, 899–926.  CrossRef CAS Web of Science Google Scholar
First citationSaeed, A., Flörke, U. & Erben, M. F. (2014). J. Sulfur Chem. 35, 318–355.  Web of Science CrossRef CAS Google Scholar
First citationSaeed, A., Larik, F. A., Jabeen, F., Mehfooz, H., Ghumro, S. A., El-Seedi, H. R., Ali, M., Channar, P. A. & Ashraf, H. (2018). Russ. J. Gen. Chem. 88, 541–550.  Web of Science CrossRef CAS Google Scholar
First citationSaeed, S., Rashid, N., Jones, P. G., Ali, M. & Hussain, R. (2010). Eur. J. Med. Chem. 45, 1323–1331.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationSelvakumaran, N., Karvembu, R., Ng, S. W. & Tiekink, E. R. T. (2011). Acta Cryst. E67, o602.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationSelvakumaran, N., Ng, S. W., Tiekink, E. R. T. & Karvembu, R. (2011). Inorg. Chim. Acta, 376, 278–284.  Web of Science CSD CrossRef CAS Google Scholar
First citationSheldrick, G. M. (1996). SADABS. University of Göttingen, Germany.  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 citationTurner, M. J., Mckinnon, J. J., Wolff, S. K., Grimwood, D. J., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). Crystal Explorer 17. The University of Western Australia.  Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWoon, D. E. & Dunning, T. H. Jr (1993). J. Chem. Phys. 98, 1358–1371.  CrossRef CAS Web of Science Google Scholar
First citationZhang, Z. & Schreiner, P. R. (2009). Chem. Soc. Rev. 38, 1187–1198.  Web of Science CrossRef PubMed CAS Google Scholar
First citationZhang, Z.-J., Zeng, Y., Jiang, Z.-Y., Shu, B.-S., Sethuraman, V. & Zhong, G.-H. (2018). Pest Manag. Sci. 74, 1736–1746.  Web of Science 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