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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structure, Hirshfeld surface and crystal void analysis, inter­molecular inter­action energies, DFT calculations and energy frameworks of 2H-benzo[b][1,4]thia­zin-3(4H)-one 1,1-dioxide

crossmark logo

aLaboratory of Organic and Physical Chemistry, Applied Bioorganic Chemistry Team, Faculty of Sciences, Ibn Zohr University, Agadir, Morocco, bUniversity of Lille, CNRS, UAR 3290, MSAP, Miniaturization for Synthesis, Analysis and Proteomics, F-59000 Lille, France, cDepartment of Physics, Hacettepe University, 06800 Beytepe, Ankara, Türkiye, dLaboratory of Applied Organic Chemistry, Faculty of Science and Technology, University of Sidi Mohamed Ben Abdellah BP 2202, Fez, Morocco, eDepartment of Chemistry, Tulane University, New Orleans, LA 70118, USA, and fLaboratory of Heterocyclic Organic Chemistry, Medicines Science Research Center, Pharmacochemistry Competence Center, Mohammed V University in Rabat, Faculté des Sciences, Av. Ibn Battouta, BP 1014, Rabat, Morocco
*Correspondence e-mail: ezaddine.irrou@edu.uiz.ac.ma

Edited by M. Weil, Vienna University of Technology, Austria (Received 13 September 2023; accepted 3 October 2023; online 19 October 2023)

This article is part of a collection of articles to commemorate the founding of the African Crystallographic Association and the 75th anniversary of the IUCr.

In the title mol­ecule, C8H7NO3S, the nitro­gen atom has a planar environment, and the thia­zine ring exhibits a screw-boat conformation. In the crystal, corrugated layers of mol­ecules parallel to the ab plane are formed by N—H⋯O and C—H⋯O hydrogen bonds together with C—H⋯π(ring) and S=O⋯π(ring) inter­actions. The layers are connected by additional C—H⋯O hydrogen bonds and π-stacking inter­actions. Hirshfeld surface analysis indicates that the most important contributions for the crystal packing are from H⋯O/O⋯H (49.4%), H⋯H (23.0%) and H⋯C/C⋯H (14.1%) inter­actions. The volume of the crystal voids and the percentage of free space were calculated as 75.4 Å3 and 9.3%. Density functional theory (DFT) computations revealed N—H⋯O and C—H⋯O hydrogen-bonding energies of 43.3, 34.7 and 34.4 kJ mol−1, respectively. Evaluation of the electrostatic, dispersion and total energy frameworks indicate that the stabilization is dominated via the electrostatic energy contribution. Moreover, the DFT-optimized structure at the B3LYP/ 6–311 G(d,p) level is compared with the experimentally determined mol­ecular structure in the solid state. The HOMO–LUMO behaviour was elucidated to determine the energy gap.

1. Chemical context

Numerous heterocyclic compounds containing sulfur and nitro­gen have been extensively studied because of their various biological applications (Gowda et al., 2011[Gowda, J., Khader, A. M. A., Kalluraya, B., Shree, P. & Shabaraya, A. R. (2011). Eur. J. Med. Chem. 46, 4100-4106.]; Sebbar et al., 2020a[Sebbar, G., Mohamed, E., Hökelek, T., Mague, J. T., Sebbar, N. K., Essassi, E. M. & Belkadi, B. (2020a). Acta Cryst. E76, 629-636.]; Fringuelli et al., 2005[Fringuelli, R., Milanese, L. & Schiaffella, F. (2005). Mini Rev. Med. Chem. 5, 1061-1073.]). In this respect, 1,4-benzo­thia­zine derivatives possess various pharmacological properties and have therapeutic applications such as anti­fungal (Kamila et al., 2006[Kamila, S., Koh, B., Zhang, H. & Biehl, E. R. (2006). Arkivoc, 2, 1-14.]), anti-inflammatory (Gowda et al., 2011[Gowda, J., Khader, A. M. A., Kalluraya, B., Shree, P. & Shabaraya, A. R. (2011). Eur. J. Med. Chem. 46, 4100-4106.]), antagonistic (Corelli et al., 1997[Corelli, F., Manetti, F., Tafi, A., Campiani, G., Nacci, V. & Botta, M. (1997). J. Med. Chem. 40, 125-131.]), anti-tumour (Abbas & Farghaly, 2010[Abbas, E. M. & Farghaly, T. A. (2010). Monatsh. Chem. 141, 661-667.]), anti­oxidant (Bakavoli et al., 2008[Bakavoli, M., Sadeghian, H., Tabatabaei, Z., Rezaei, E., Rahimizadeh, M. & Nikpour, M. (2008). J. Mol. Model. 14, 471-478.]), anti­pyretic (Warren & Knaus, 1987[Warren, B. K. & Knaus, E. E. (1987). Eur. J. Med. Chem. 22, 411-415.]), anti­hypertensive (Fringuelli et al., 2005[Fringuelli, R., Milanese, L. & Schiaffella, F. (2005). Mini Rev. Med. Chem. 5, 1061-1073.]) or anti­bacterial effects (Sebbar et al., 2016[Sebbar, N. K., Mekhzoum, M. E. M., Essassi, E. M., Zerzouf, A., Talbaoui, A., Bakri, Y., Saadi, M. & Ammari, L. E. (2016). Res. Chem. Intermed. 42, 6845-6862.], 2020a[Sebbar, G., Mohamed, E., Hökelek, T., Mague, J. T., Sebbar, N. K., Essassi, E. M. & Belkadi, B. (2020a). Acta Cryst. E76, 629-636.]).

Continuing our research on the development of new 1,4-benzo­thia­zine derivatives with potential pharmacological applications, we carried out the oxidation of 3,4-di­hydro-2H-1,4-benzo­thia­zin-3-one by potassium permanganate in order to obtain 2H-benzo[b][1,4]thia­zin-3(4H)-one 1,1-dioxide (I)[link] with good yield. We report herein the mol­ecular and crystal structure of this compound, as well as Hirshfeld surface analysis and DFT-computational studies carried out at the B3LYP/6–31 G(d,p) and B3LYP/6–311 G(d,p) levels.

[Scheme 1]

2. Structural commentary

A puckering analysis (Cremer & Pople, 1975[Cremer, D. & Pople, J. A. (1975). J. Am. Chem. Soc. 97, 1354-1358.]) of the thia­zine ring (C1, C6, N1, C7, C8, S1) gave the parameters Q = 0.5138 (6) Å, θ = 60.49 (7)° and φ = 326.72 (8)°. The distorted screw-boat conformation places O2 in an axial position and O3 in a pseudo-equatorial position (Fig. 1[link]). The angles about N1 sum up to 360° within experimental error, indicating involvement of the lone pair in the C–N bond. This is reflected in the N1—C7 and N1—C6 distances of 1.3661 (9) and 1.4043 (9) Å, respectively.

[Figure 1]
Figure 1
The title mol­ecule with atom labelling and displacement ellipsoids drawn at the 50% probability level.

3. Supra­molecular features

In the crystal, N1—H1⋯O3 hydrogen bonds (Table 1[link]) form chains of mol­ecules extending parallel to the a axis. These chains are connected into corrugated layers parallel to the ab plane by C8—H8B⋯O2 hydrogen bonds together with C8—H8ACg2 and S1=O2⋯Cg2i inter­actions [O2⋯Cg2 = 3.6233 (7) Å, S1⋯Cg2 = 4.1655 (5) Å, S1=O2⋯Cg2i = 101.77 (3)°; symmetry code: (i) −x + [{1\over 2}], y + [{1\over 2}], −z + [{3\over 2}]; Fig. 2[link]]. The layers are connected by C5—H5⋯O1 hydrogen bonds (Table 1[link]) and slipped ππ stacking inter­actions between inversion-related C1–C6 rings [Cg2⋯Cg2(1 −x, 1 −y, 1 −z) = 3.7353 (5) Å, slippage = 1.55 Å] into a tri-periodic network structure (Fig. 3[link]).

Table 1
Hydrogen-bond geometry (Å, °)

Cg2 is the centroid of the C1–C6 benzene ring.

D—H⋯A D—H H⋯A DA D—H⋯A
N1—H1⋯O3i 0.89 (1) 2.08 (1) 2.9472 (8) 165 (1)
C5—H5⋯O1ii 0.95 2.58 3.2330 (9) 126
C8—H8ACg2iii 0.99 2.93 3.7933 (8) 146
C8—H8B⋯O2iv 0.99 2.39 3.2126 (9) 140
Symmetry codes: (i) [x+1, y, z]; (ii) [-x+{\script{3\over 2}}, y-{\script{1\over 2}}, -z+{\script{3\over 2}}]; (iii) [-x+{\script{1\over 2}}, y+{\script{1\over 2}}, -z+{\script{3\over 2}}]; (iv) [-x+{\script{1\over 2}}, y-{\script{1\over 2}}, -z+{\script{3\over 2}}].
[Figure 2]
Figure 2
The crystal structure of (I)[link] viewed along the c axis with N—H⋯O and C—H⋯O hydrogen bonds depicted, respectively, by violet and black dashed lines. C—H⋯π(ring) and C=O⋯π(ring) inter­actions are depicted, respectively, by green and dark-pink dashed lines and non-inter­acting hydrogen atoms are omitted for clarity.
[Figure 3]
Figure 3
The crystal structure of (I)[link] viewed along the b axis with N—H⋯O and C—H⋯O hydrogen bonds depicted, respectively, by violet and black dashed lines. C—H⋯π(ring), C=O⋯π(ring) and slipped π-stacking inter­actions are depicted, respectively, by green, dark-pink and orange dashed lines. Non-inter­acting hydrogen atoms are omitted for clarity.

4. Hirshfeld surface analysis

To visualize the inter­molecular inter­actions in the crystal of (I)[link], a Hirshfeld surface (HS) analysis (Hirshfeld, 1977[Hirshfeld, H. L. (1977). Theor. Chim. Acta, 44, 129-138.]) was carried out with Crystal Explorer (Spackman et al., 2021[Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006-1011.]). In the HS plotted over dnorm in the range −0.4976 to 1.2253 a.u. (Fig. 4[link]), the white surface indicates contacts with distances equal to the sum of van der Waals radii and the red and blue colours indicate distances shorter (in close contact) or longer (distant contact) than the van der Waals radii, respectively (Venkatesan et al., 2016[Venkatesan, P., Thamotharan, S., Ilangovan, A., Liang, H. & Sundius, T. (2016). Spectrochim. Acta A Mol. Biomol. Spectrosc. 153, 625-636.]). The bright-red spots indicate their roles as the respective donors and/or acceptors; they also appear as blue and red regions corresponding to positive and negative potentials on the HS mapped over electrostatic potential (Spackman et al., 2008[Spackman, M. A., McKinnon, J. J. & Jayatilaka, D. (2008). CrystEngComm, 10, 377-388.]; Jayatilaka et al., 2005[Jayatilaka, D., Grimwood, D. J., Lee, A., Lemay, A., Russel, A. J., Taylor, C., Wolff, S. K., Cassam-Chenai, P. & Whitton, A. (2005). TONTO - A System for Computational Chemistry. Available at: http://hirshfeldsurface.net/]) in the range −0.05 to 0.05 a.u., as shown in Fig. 5[link]. The blue regions indicate positive electrostatic potential (hydrogen-bond donors), while the red regions indicate negative electrostatic potential (hydrogen-bond acceptors). The shape-index of the HS is a tool to visualize the ππ stacking by the presence of adjacent red and blue triangles. Fig. 6[link] clearly suggests that there are ππ inter­actions in (I)[link]. The overall two-dimensional fingerprint plot, Fig. 7[link]a, and those delineated into H⋯O/O⋯H, H⋯H, H⋯C/C⋯H, C⋯O/O⋯C,C⋯C, H⋯N/N⋯H, O⋯O and C⋯N/N⋯C contacts (McKinnon et al., 2007[McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. 3814-3816.]) are illustrated in Fig. 7[link]bi, respectively, together with their relative contributions to the Hirshfeld surface. The most important inter­action is H⋯O/O⋯H, contributing 49.4% to the overall crystal packing, which is reflected in Fig. 7[link]b, where the symmetric pair of spikes is observed with the tips at de + di = 1.98 Å. The H⋯H contacts contribute 23.0% to the overall crystal packing, which is reflected in Fig. 7[link]c as widely scattered points of high density due to the large hydrogen content of the mol­ecule with the tip at de = di = 1.13 Å. In the presence of C—H⋯π inter­actions, the pair of characteristic wings in the fingerprint plot delineated into H⋯C/C⋯H contacts, Fig. 7[link]d, make a 14.1% contribution to the HS and viewed with the tips at de + di = 2.59 Å. The wing pair of C⋯O/O⋯C contacts (Fig. 7[link]e) with 4.9% contribution to the HS is viewed at de + di = 3.30 Å. The C⋯C contacts (Fig. 7[link]f) appearing as a bullet-shaped distribution of points make a contribution of 3.7% to the HS with the tip at de = di = 1.70 Å. The spikes of H⋯N/N⋯H contacts (Fig. 7[link]g) with 3.2% contribution to the HS are viewed at de + di = 2.75 Å. Finally, the O⋯O (Fig. 7[link]h) and C⋯N/N⋯C (Fig. 7[link]i) contacts contribute 1.3% and 0.4%, respectively, to the HS. The Hirshfeld surface representations with the function dnorm plotted onto the surface are shown for the H⋯O/O⋯H, H⋯H and H⋯C/C⋯H inter­actions in Fig. 8[link]ac, respectively. The Hirshfeld surface analysis confirms the importance of H-atom contacts in establishing the packing. The large number of H⋯O/O⋯H, H⋯H and H⋯C/C⋯H inter­actions suggest that van der Waals inter­actions play the major role in the crystal packing (Hathwar et al., 2015[Hathwar, V. R., Sist, M., Jørgensen, M. R. V., Mamakhel, A. H., Wang, X., Hoffmann, C. M., Sugimoto, K., Overgaard, J. & Iversen, B. B. (2015). IUCrJ, 2, 563-574.]).

[Figure 4]
Figure 4
View of the three-dimensional Hirshfeld surface of (I)[link] plotted over dnorm.
[Figure 5]
Figure 5
View of the three-dimensional Hirshfeld surface of (I)[link] plotted over electrostatic potential energy using the STO-3 G basis set at the Hartree–Fock level of theory.
[Figure 6]
Figure 6
Hirshfeld surface of (I)[link] plotted over shape-index.
[Figure 7]
Figure 7
The full two-dimensional fingerprint plots for the title compound, showing (a) all inter­actions, (b) H⋯O/O⋯H, (c) H⋯H, (d) H⋯C/C⋯H, (e) O⋯C/C⋯O, (f) C⋯C, (g) H⋯N/N⋯H, (h) O⋯O and (i) C⋯N/N⋯C inter­actions. The di and de values are the closest inter­nal and external distances (in Å) from given points on the Hirshfeld surface contacts.
[Figure 8]
Figure 8
Hirshfeld surface representations of (I)[link] with the function dnorm plotted onto the surface for (a) H⋯O/O⋯H, (b) H⋯H and (c) H⋯C/C⋯H inter­actions.

The strength of the crystal packing is important for determining the response to an applied mechanical force. If the crystal packing results in significant voids, then the mol­ecules are not tightly packed and a small amount of applied external mechanical force may easily break the crystal. To check the mechanical stability of the crystal, a void analysis was performed by adding up the electron densities of the spherically symmetric atoms contained in the asymmetric unit (Turner et al., 2011[Turner, M. J., McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2011). CrystEngComm, 13, 1804-1813.]). The void surface is defined as an isosurface of the procrystal electron density and is calculated for the whole unit cell where the void surface meets the boundary of the unit cell and capping faces are generated to create an enclosed volume. The volume of the crystal voids (Fig. 9[link]a,b) and the percentage of free space in the unit cell are calculated as 75.4 Å3 and 9.3%, respectively. Thus, the crystal packing appears compact and the mechanical stability should be substantial.

[Figure 9]
Figure 9
Graphical views of voids in the crystal packing of (I)[link], (a) along the a axis and (b) along the b axis.

5. Inter­action energy calculations and energy frameworks

The inter­molecular inter­action energies were calculated using the CEB3LYP/631G(d,p) energy model available in CrystalExplorer (Spackman et al., 2021[Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006-1011.]), where a cluster of mol­ecules is generated by applying crystallographic symmetry operations with respect to a selected central mol­ecule within the radius of 3.8 Å by default (Turner et al., 2014[Turner, M. J., Grabowsky, S., Jayatilaka, D. & Spackman, M. A. (2014). J. Phys. Chem. Lett. 5, 4249-4255.]). The total inter­molecular energy (Etot) is the sum of electrostatic (Eele), polarization (Epol), dispersion (Edis) and exchange-repulsion (Erep) energies (Turner et al., 2015[Turner, M. J., Thomas, S. P., Shi, M. W., Jayatilaka, D. & Spackman, M. A. (2015). Chem. Commun. 51, 3735-3738.]) with scale factors of 1.057, 0.740, 0.871 and 0.618, respectively (Mackenzie et al., 2017[Mackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). IUCrJ, 4, 575-587.]). Hydrogen-bonding inter­action energies (in kJ mol−1) were calculated to be [−18.5 (Eele), −5.2 (Epol), −41.4 (Edis), 26.2 (Erep) and −43.3 (Etot)] for N1—H1⋯O3, [−22.4 (Eele), −4.8 (Epol), −28.3 (Edis), 27.9 (Erep) and −34.7 (Etot)] for C8—H8B⋯O2 and [−20.6 (Eele), −5.8 (Epol), −24.6 (Edis), 21.2 (Erep) and −34.4 (Etot)] for C5—H5⋯O1.

Energy frameworks combine the calculation of inter­molecular inter­action energies with a graphical representation of their magnitude (Turner et al., 2015[Turner, M. J., Thomas, S. P., Shi, M. W., Jayatilaka, D. & Spackman, M. A. (2015). Chem. Commun. 51, 3735-3738.]). Energies between mol­ecular pairs are represented as cylinders joining the centroids of pairs of mol­ecules with the cylinder radius proportional to the relative strength of the corresponding inter­action energy. Energy frameworks were constructed for Eele (shown in Fig. 10[link]), Edis and Etot. The evaluation of the electrostatic, dispersion and total energy frameworks indicate that the stabilization is dominated via the electrostatic energy contribution in the crystal structure of (I)[link].

[Figure 10]
Figure 10
The energy framework for the electrostatic energy, viewed down the b axis for a cluster of mol­ecules, where the a axis is vertical and the c axis is horizontal. The cylindrical radius is proportional to the relative strength of the corresponding energy and adjusted to the scale factor of 80 with a cut-off value of 5 kJ mol−1 within 2 × 2 × 2 unit cells.

6. DFT calculations

The optimized structure of (I)[link] was computed in the gas phase using density functional theory (DFT) with the standard B3LYP functional and 6–311 G(d,p) basis-set calculations (Becke, 1993[Becke, A. D. (1993). J. Chem. Phys. 98, 5648-5652.]), employing the GAUSSIAN 09 software (Frisch et al., 2009[Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, O., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2009). GAUSSIAN09. Gaussian Inc., Wallingford, CT, US]). The theoretical and experimental results exhibit a good agreement, as summarized in Table 2[link].

Table 2
Comparison of the selected (X-ray and DFT) geometric data (Å, °)

Bonds/angles X-ray B3LYP/6–311G(d,p)
S1—O2 1.4450 (6) 1.50996
S1—O3 1.4464 (6) 1.59088
S1—C1 1.7478 (6) 1.78874
S1—C8 1.7649 (7) 1.80529
O1—C7 1.2203 (8) 1.21656
N1—C7 1.3661 (9) 1.37417
N1—C6 1.4043 (9) 1.39867
O2—S1—O3 117.64 (4) 118.09813
O2—S1—C1 109.21 (3) 109.40063
O3—S1—C1 109.56 (3) 109.84640
O2—S1—C8 108.59 (3) 109.10007
O3—S1—C8 109.59 (3) 109.62080
C1—S1—C8 100.95 (3) 99.96775
C7—N1—C6 127.24 (6) 127.88849
C7—N1—H1 116.1 (10) 115.98354

The highest-occupied mol­ecular orbital (HOMO), functioning as an electron donor, and the lowest-unoccupied mol­ecular orbital (LUMO), acting as an electron acceptor, serve as vital parameters in quantum chemistry. A small energy gap signifies high mol­ecular polarizability and enhanced chemical reactivity. The DFT calculations provided crucial insights into the reactivity and site selectivity of the mol­ecular framework. Parameters such as EHOMO and ELUMO, electronegativity (χ), hardness (η), dipole moment (μ), electrophilicity (ω) and softness (σ) are compiled in Table 3[link]. Both η and σ are essential for assessing reactivity and stability. The electron transition from HOMO to LUMO energy levels is depicted in Fig. 11[link]. Notably, both HOMO and LUMO are localized within the plane spanning the entire 2H-benzo[b][1,4]thia­zin-3(4H)-one 1,1-dioxide ring. The energy band gap [ΔE = ELUMO − EHOMO] for the mol­ecule is 11.7261 eV, and the energies of the frontier mol­ecular orbitals, EHOMO and ELUMO, are −9.6740 eV and 2.0522 eV, respectively.

Table 3
Calculated energies

Mol­ecular Energy (a.u.) (eV) Compound (I)
Total Energy, TE (eV) −26615,8936
EHOMO (eV) −9.6740
ELUMO (eV) 2.0522
Gap, ΔE (eV) 11.7261
Dipole moment, μ (Debye) 7.583751
Ionization potential, I (eV) 9.6740
Electron affinity, A 2.0522
Electronegativity, χ −3.8109
Hardness, η −5.8631
Electrophilicity index, ω −1.2385
Softness σ, −0.1706
Fraction of electron transferred, ΔN −0.9219
[Figure 11]
Figure 11
The energy band gap of (I)[link].

7. Database survey

A search in the Cambridge Structural Database (CSD, updated March 2023; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) for compounds containing the fragment II (R1 = Ph or 2-ClC6H4, R2 = C; Fig. 12[link]), gave 14 hits. With R1 = Ph, and with R2 = CH2COOCH2CH3 (IIa; Sebbar et al., 2020b[Sebbar, N. K., Labd, T. M., Ellouz, M., Essassi, E. M., Zerzouf, A., Karrouchi, K., Ouzidan, Y., Mennane, Z. & Mague, J. T. (2020b). Iran. J. Chem. Chem. Eng. 39, 53-67.]), CH2COOH (IIb; Sebbar et al., 2016[Sebbar, N. K., Mekhzoum, M. E. M., Essassi, E. M., Zerzouf, A., Talbaoui, A., Bakri, Y., Saadi, M. & Ammari, L. E. (2016). Res. Chem. Intermed. 42, 6845-6862.]), CH2C≡CH (IIc; Sebbar et al., 2014[Sebbar, N. K., Zerzouf, A., Essassi, E. M., Saadi, M. & El Ammari, L. (2014). Acta Cryst. E70, o614.]) and C5H8NO2 (IId; Sebbar et al., 2016[Sebbar, N. K., Mekhzoum, M. E. M., Essassi, E. M., Zerzouf, A., Talbaoui, A., Bakri, Y., Saadi, M. & Ammari, L. E. (2016). Res. Chem. Intermed. 42, 6845-6862.]) (Fig. 12[link]) are matching candidates. Other examples with R1 = 4-FC6H4 and R2 = CH2C≡CH (Hni et al., 2019[Hni, B., Sebbar, N. K., Hökelek, T., Ouzidan, Y., Moussaif, A., Mague, J. T. & Essassi, E. M. (2019). Acta Cryst. E75, 372-377.]) and R1 = 2-ClC6H4, R2 = CH2C≡CH (Sebbar et al., 2017[Sebbar, N. K., Ellouz, M., Ouzidan, Y., Kaur, M., Essassi, E. M. & Jasinski, J. P. (2017). IUCrData, 2, x170889.]) are also known.

[Figure 12]
Figure 12
The mol­ecular moieties (II) used for the CSD database search.

8. Synthesis and crystallization

3,4-Di­hydro-2H-1,4-benzo­thia­zin-3-one (1.2 mmol) was dissolved in 3 ml of acetic acid and added dropwise into a solution of potassium permanganate (1.81 mmol) in 6 ml of water. After stirring for one h at room temperature, a solution of sodium thio­sulfate penta­hydrate (20%wt) was added to react with excessive potassium permanganate. The precipitate obtained was filtered and recrystallized from ethanol to yield single-crystals suitable for X-ray structure analysis..

9. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 4[link]. H-atoms attached to carbon were placed in calculated positions (C—H = 0.95–0.99 Å) and were included as riding contributions with isotropic displacement parameters 1.2 or 1.5 times those of the attached atoms. That attached to nitro­gen was placed in a location derived from a difference map and refined with a DFIX 0.91 0.01 instruction. Two reflections affected by the beamstop were omitted from the final refinement.

Table 4
Experimental details

Crystal data
Chemical formula C8H7NO3S
Mr 197.21
Crystal system, space group Monoclinic, P21/n
Temperature (K) 125
a, b, c (Å) 7.2179 (6), 9.5043 (8), 11.9945 (9)
β (°) 97.584 (2)
V3) 815.64 (11)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.37
Crystal size (mm) 0.39 × 0.21 × 0.16
 
Data collection
Diffractometer Bruker D8 QUEST PHOTON 3 diffractometer
Absorption correction Multi-scan (SADABS; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.])
Tmin, Tmax 0.91, 0.94
No. of measured, independent and observed [I > 2σ(I)] reflections 47688, 3957, 3739
Rint 0.027
(sin θ/λ)max−1) 0.836
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.025, 0.074, 1.04
No. of reflections 3957
No. of parameters 122
No. of restraints 1
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.52, −0.36
Computer programs: APEX4 and SAINT (Bruker, 2021[Bruker (2021). APEX4 and SAINT. Bruker AXS LLC, Madison, Wisconsin, USA.]), SHELXT (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2018/1 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), DIAMOND (Brandenburg & Putz, 2012[Brandenburg, K. & Putz, H. (2012). DIAMOND. Crystal Impact GbR, Bonn, Germany.]) and SHELXTL (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]).

Supporting information


Computing details top

2H-Benzo[b][1,4]thiazin-3(4H)-one 1,1-dioxide top
Crystal data top
C8H7NO3SF(000) = 408
Mr = 197.21Dx = 1.606 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 7.2179 (6) ÅCell parameters from 9215 reflections
b = 9.5043 (8) Åθ = 2.9–36.4°
c = 11.9945 (9) ŵ = 0.37 mm1
β = 97.584 (2)°T = 125 K
V = 815.64 (11) Å3Prism, colourless
Z = 40.39 × 0.21 × 0.16 mm
Data collection top
Bruker D8 QUEST PHOTON 3
diffractometer
3957 independent reflections
Radiation source: fine-focus sealed tube3739 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.027
Detector resolution: 7.3910 pixels mm-1θmax = 36.4°, θmin = 3.6°
φ and ω scansh = 1212
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
k = 1515
Tmin = 0.91, Tmax = 0.94l = 1920
47688 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)] = 0.025Hydrogen site location: mixed
wR(F2) = 0.074H atoms treated by a mixture of independent and constrained refinement
S = 1.04 w = 1/[σ2(Fo2) + (0.0383P)2 + 0.2724P]
where P = (Fo2 + 2Fc2)/3
3957 reflections(Δ/σ)max = 0.001
122 parametersΔρmax = 0.52 e Å3
1 restraintΔρmin = 0.36 e Å3
Special details top

Experimental. The diffraction data were obtained from 9 sets of frames, each of width 0.5° in ω or φ, collected with scan parameters determined by the "strategy" routine in APEX4. The scan time was 15 sec/frame.

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.

Refinement. Refinement of F2 against ALL reflections. The weighted R-factor wR and goodness of fit S are based on F2, conventional R-factors R are based on F, with F set to zero for negative F2. The threshold expression of F2 > 2sigma(F2) is used only for calculating R-factors(gt) etc. and is not relevant to the choice of reflections for refinement. R-factors based on F2 are statistically about twice as large as those based on F, and R- factors based on ALL data will be even larger.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
S10.16068 (2)0.52143 (2)0.71754 (2)0.01099 (4)
O10.61228 (8)0.59530 (6)0.90859 (5)0.01991 (10)
O20.16268 (8)0.66721 (6)0.68352 (5)0.01797 (10)
O30.01882 (7)0.45794 (7)0.72667 (5)0.01832 (10)
N10.57806 (8)0.47975 (6)0.74211 (5)0.01411 (10)
H10.7001 (12)0.4900 (15)0.7398 (13)0.027 (3)*
C10.27959 (8)0.42044 (6)0.62800 (5)0.01059 (9)
C20.17997 (9)0.35615 (7)0.53412 (5)0.01342 (10)
H20.0473970.3612370.5217760.016*
C30.27651 (11)0.28460 (7)0.45882 (6)0.01663 (11)
H30.2108470.2427270.3932920.020*
C40.47092 (11)0.27481 (8)0.48036 (6)0.01841 (12)
H40.5368250.2248370.4293800.022*
C50.56985 (10)0.33681 (8)0.57499 (6)0.01630 (11)
H50.7019380.3276050.5889420.020*
C60.47486 (8)0.41274 (7)0.64966 (5)0.01164 (10)
C70.51172 (9)0.53288 (7)0.83499 (5)0.01315 (10)
C80.30863 (9)0.50214 (7)0.84603 (5)0.01299 (10)
H8A0.2665780.5667510.9023740.016*
H8B0.2981670.4047530.8738250.016*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
S10.00776 (7)0.01295 (7)0.01219 (7)0.00109 (4)0.00109 (4)0.00153 (4)
O10.0173 (2)0.0236 (3)0.0174 (2)0.00682 (19)0.00298 (17)0.00209 (18)
O20.0205 (2)0.0129 (2)0.0200 (2)0.00460 (17)0.00061 (18)0.00015 (16)
O30.00783 (18)0.0262 (3)0.0214 (2)0.00199 (17)0.00373 (16)0.00439 (19)
N10.0078 (2)0.0195 (2)0.0149 (2)0.00112 (17)0.00092 (16)0.00008 (18)
C10.0092 (2)0.0114 (2)0.0113 (2)0.00061 (16)0.00190 (16)0.00003 (16)
C20.0138 (2)0.0137 (2)0.0125 (2)0.00096 (19)0.00073 (18)0.00101 (18)
C30.0212 (3)0.0150 (3)0.0140 (2)0.0010 (2)0.0038 (2)0.00282 (19)
C40.0218 (3)0.0166 (3)0.0184 (3)0.0025 (2)0.0088 (2)0.0024 (2)
C50.0133 (2)0.0177 (3)0.0190 (3)0.0033 (2)0.0064 (2)0.0004 (2)
C60.0093 (2)0.0130 (2)0.0130 (2)0.00103 (17)0.00242 (17)0.00133 (17)
C70.0114 (2)0.0143 (2)0.0133 (2)0.00146 (18)0.00035 (18)0.00154 (18)
C80.0115 (2)0.0162 (2)0.0113 (2)0.00114 (19)0.00154 (18)0.00122 (18)
Geometric parameters (Å, º) top
S1—O21.4450 (6)C2—H20.9500
S1—O31.4464 (6)C3—C41.3960 (11)
S1—C11.7478 (6)C3—H30.9500
S1—C81.7649 (7)C4—C51.3898 (11)
O1—C71.2203 (8)C4—H40.9500
N1—C71.3661 (9)C5—C61.3982 (9)
N1—C61.4043 (9)C5—H50.9500
N1—H10.890 (8)C7—C81.5175 (9)
C1—C21.3945 (9)C8—H8A0.9900
C1—C61.4010 (9)C8—H8B0.9900
C2—C31.3892 (10)
O2—S1—O3117.64 (4)C5—C4—C3121.22 (6)
O2—S1—C1109.21 (3)C5—C4—H4119.4
O3—S1—C1109.56 (3)C3—C4—H4119.4
O2—S1—C8108.59 (3)C4—C5—C6119.96 (6)
O3—S1—C8109.59 (3)C4—C5—H5120.0
C1—S1—C8100.95 (3)C6—C5—H5120.0
C7—N1—C6127.24 (6)C5—C6—C1118.33 (6)
C7—N1—H1116.1 (10)C5—C6—N1119.08 (6)
C6—N1—H1116.7 (10)C1—C6—N1122.58 (6)
C2—C1—C6121.72 (6)O1—C7—N1122.05 (6)
C2—C1—S1119.61 (5)O1—C7—C8121.31 (6)
C6—C1—S1118.58 (5)N1—C7—C8116.53 (6)
C3—C2—C1119.34 (6)C7—C8—S1112.57 (4)
C3—C2—H2120.3C7—C8—H8A109.1
C1—C2—H2120.3S1—C8—H8A109.1
C2—C3—C4119.38 (6)C7—C8—H8B109.1
C2—C3—H3120.3S1—C8—H8B109.1
C4—C3—H3120.3H8A—C8—H8B107.8
O2—S1—C1—C293.69 (6)C2—C1—C6—C50.74 (9)
O3—S1—C1—C236.48 (6)S1—C1—C6—C5177.28 (5)
C8—S1—C1—C2152.03 (5)C2—C1—C6—N1178.32 (6)
O2—S1—C1—C682.92 (6)S1—C1—C6—N11.78 (8)
O3—S1—C1—C6146.91 (5)C7—N1—C6—C5165.40 (7)
C8—S1—C1—C631.36 (6)C7—N1—C6—C115.55 (10)
C6—C1—C2—C31.16 (10)C6—N1—C7—O1176.50 (7)
S1—C1—C2—C3175.34 (5)C6—N1—C7—C87.37 (10)
C1—C2—C3—C41.92 (10)O1—C7—C8—S1141.26 (6)
C2—C3—C4—C50.81 (11)N1—C7—C8—S142.58 (7)
C3—C4—C5—C61.12 (11)O2—S1—C8—C764.45 (5)
C4—C5—C6—C11.86 (10)O3—S1—C8—C7165.81 (5)
C4—C5—C6—N1177.23 (6)C1—S1—C8—C750.29 (5)
Hydrogen-bond geometry (Å, º) top
Cg2 is the centroid of the C1–C6 benzene ring.
D—H···AD—HH···AD···AD—H···A
N1—H1···O3i0.89 (1)2.08 (1)2.9472 (8)165 (1)
C5—H5···O1ii0.952.583.2330 (9)126
C8—H8A···Cg2iii0.992.933.7933 (8)146
C8—H8B···O2iv0.992.393.2126 (9)140
Symmetry codes: (i) x+1, y, z; (ii) x+3/2, y1/2, z+3/2; (iii) x+1/2, y+1/2, z+3/2; (iv) x+1/2, y1/2, z+3/2.
Comparison of the selected (X-ray and DFT) geometric data (Å, °) top
Bonds/anglesX-rayB3LYP/6-311G(d,p)
S1—O21.4450 (6)1.50996
S1—O31.4464 (6)1.59088
S1—C11.7478 (6)1.78874
S1—C81.7649 (7)1.80529
O1—C71.2203 (8)1.21656
N1—C71.3661 (9)1.37417
N1—C61.4043 (9)1.39867
O2—S1—O3117.64 (4)118.09813
O2—S1—C1109.21 (3)109.40063
O3—S1—C1109.56 (3)109.84640
O2—S1—C8108.59 (3)109.10007
O3—S1—C8109.59 (3)109.62080
C1—S1—C8100.95 (3)99.96775
C7—N1—C6127.24 (6)127.88849
C7—N1—H1116.1 (10)115.98354
Calculated energies top
Molecular Energy (a.u.) (eV)Compound (I)
Total Energy, TE (eV)-26615,8936
EHOMO (eV)-9.6740
ELUMO (eV)2.0522
Gap, ΔE (eV)11.7261
Dipole moment, µ (Debye)7.583751
Ionization potential, I (eV)9.6740
Electron affinity, A2.0522
Electronegativity, χ-3.8109
Hardness, η-5.8631
Electrophilicity index, ω-1.2385
Softness σ,-0.1706
Fraction of electron transferred, ΔN-0.9219
 

Funding information

JTM thanks Tulane University for support of the Tulane Crystallography Laboratory. TH is grateful to Hacettepe University Scientific Research Project Unit (grant No. 013 D04 602 004).

References

First citationAbbas, E. M. & Farghaly, T. A. (2010). Monatsh. Chem. 141, 661–667.  CrossRef CAS Google Scholar
First citationBakavoli, M., Sadeghian, H., Tabatabaei, Z., Rezaei, E., Rahimizadeh, M. & Nikpour, M. (2008). J. Mol. Model. 14, 471–478.  CrossRef PubMed CAS Google Scholar
First citationBecke, A. D. (1993). J. Chem. Phys. 98, 5648–5652.  CrossRef CAS Web of Science Google Scholar
First citationBrandenburg, K. & Putz, H. (2012). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBruker (2021). APEX4 and SAINT. Bruker AXS LLC, Madison, Wisconsin, USA.  Google Scholar
First citationCorelli, F., Manetti, F., Tafi, A., Campiani, G., Nacci, V. & Botta, M. (1997). J. Med. Chem. 40, 125–131.  CrossRef CAS PubMed Google Scholar
First citationCremer, D. & Pople, J. A. (1975). J. Am. Chem. Soc. 97, 1354–1358.  CrossRef CAS Web of Science Google Scholar
First citationFringuelli, R., Milanese, L. & Schiaffella, F. (2005). Mini Rev. Med. Chem. 5, 1061–1073.  Web of Science CrossRef PubMed CAS Google Scholar
First citationFrisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, O., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2009). GAUSSIAN09. Gaussian Inc., Wallingford, CT, US  Google Scholar
First citationGowda, J., Khader, A. M. A., Kalluraya, B., Shree, P. & Shabaraya, A. R. (2011). Eur. J. Med. Chem. 46, 4100–4106.  Web of Science CrossRef CAS PubMed 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 citationHathwar, V. R., Sist, M., Jørgensen, M. R. V., Mamakhel, A. H., Wang, X., Hoffmann, C. M., Sugimoto, K., Overgaard, J. & Iversen, B. B. (2015). IUCrJ, 2, 563–574.  Web of Science CSD CrossRef CAS PubMed IUCr Journals Google Scholar
First citationHirshfeld, H. L. (1977). Theor. Chim. Acta, 44, 129–138.  CrossRef CAS Web of Science Google Scholar
First citationHni, B., Sebbar, N. K., Hökelek, T., Ouzidan, Y., Moussaif, A., Mague, J. T. & Essassi, E. M. (2019). Acta Cryst. E75, 372–377.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationJayatilaka, D., Grimwood, D. J., Lee, A., Lemay, A., Russel, A. J., Taylor, C., Wolff, S. K., Cassam-Chenai, P. & Whitton, A. (2005). TONTO - A System for Computational Chemistry. Available at: http://hirshfeldsurface.net/  Google Scholar
First citationKamila, S., Koh, B., Zhang, H. & Biehl, E. R. (2006). Arkivoc, 2, 1–14.  Google Scholar
First citationKrause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.  Web of Science CSD CrossRef ICSD CAS IUCr Journals Google Scholar
First citationMackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. (2017). IUCrJ, 4, 575–587.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationMcKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2007). Chem. Commun. 3814–3816.  Google Scholar
First citationSebbar, G., Mohamed, E., Hökelek, T., Mague, J. T., Sebbar, N. K., Essassi, E. M. & Belkadi, B. (2020a). Acta Cryst. E76, 629–636.  CSD CrossRef IUCr Journals Google Scholar
First citationSebbar, N. K., Ellouz, M., Ouzidan, Y., Kaur, M., Essassi, E. M. & Jasinski, J. P. (2017). IUCrData, 2, x170889.  Google Scholar
First citationSebbar, N. K., Labd, T. M., Ellouz, M., Essassi, E. M., Zerzouf, A., Karrouchi, K., Ouzidan, Y., Mennane, Z. & Mague, J. T. (2020b). Iran. J. Chem. Chem. Eng. 39, 53–67.  Google Scholar
First citationSebbar, N. K., Mekhzoum, M. E. M., Essassi, E. M., Zerzouf, A., Talbaoui, A., Bakri, Y., Saadi, M. & Ammari, L. E. (2016). Res. Chem. Intermed. 42, 6845–6862.  Web of Science CSD CrossRef CAS Google Scholar
First citationSebbar, N. K., Zerzouf, A., Essassi, E. M., Saadi, M. & El Ammari, L. (2014). Acta Cryst. E70, o614.  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. (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., McKinnon, J. J. & Jayatilaka, D. (2008). CrystEngComm, 10, 377–388.  CAS Google Scholar
First citationSpackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006–1011.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationTurner, M. J., Grabowsky, S., Jayatilaka, D. & Spackman, M. A. (2014). J. Phys. Chem. Lett. 5, 4249–4255.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTurner, M. J., McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2011). CrystEngComm, 13, 1804–1813.  Web of Science CrossRef CAS Google Scholar
First citationTurner, M. J., Thomas, S. P., Shi, M. W., Jayatilaka, D. & Spackman, M. A. (2015). Chem. Commun. 51, 3735–3738.  Web of Science CrossRef CAS Google Scholar
First citationVenkatesan, P., Thamotharan, S., Ilangovan, A., Liang, H. & Sundius, T. (2016). Spectrochim. Acta A Mol. Biomol. Spectrosc. 153, 625–636.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationWarren, B. K. & Knaus, E. E. (1987). Eur. J. Med. Chem. 22, 411–415.  CrossRef CAS Web of Science 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