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

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296

Synthesis, crystal structure and in-silico evaluation of aryl­sul­fon­amide Schiff bases for potential activity against colon cancer

crossmark logo

aDepartment of Chemistry, University of Lagos, Akoka-Yaba, Lagos, Nigeria, and bDepartment of Chemistry, Nelson Mandela University, Port Elizabeth 6031, South Africa
*Correspondence e-mail: adeniyi.ogunlaja@mandela.ac.za, familonio@unilag.edu.ng

Edited by A. Lemmerer, University of the Witwatersrand, South Africa (Received 5 December 2023; accepted 11 March 2024; online 28 March 2024)

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.

This report presents a comprehensive investigation into the synthesis and characterization of Schiff base com­pounds derived from benzene­sul­fon­amide. The synthesis process, involved the reaction between N-cyclo­amino-2-sulf­anil­amide and various substituted o-salicyl­aldehydes, resulted in a set of com­pounds that were subjected to rigorous characterization using advanced spectral techniques, including 1H NMR, 13C NMR and FT–IR spectroscopy, and single-crystal X-ray diffraction. Furthermore, an in-depth assessment of the synthesized com­pounds was conducted through Absorption, Distribution, Metabolism, Excretion and Toxicity (ADMET) analysis, in conjunction with docking studies, to elucidate their pharmacokinetic profiles and potential. Impressively, the ADMET analysis showcased encouraging drug-likeness properties of the newly synthesized Schiff bases. These computational findings were substanti­ated by mol­ecular properties derived from density functional theory (DFT) calculations using the B3LYP/6-31G* method within the Jaguar Module of Schrödinger 2023-2 from Maestro (Schrodinger LLC, New York, USA). The ex­plor­ation of frontier mol­ecular orbitals (HOMO and LUMO) enabled the computation of global reactivity descriptors (GRDs), encompassing charge separation (Egap) and global softness (S). Notably, within this analysis, one Schiff base, namely, 4-bromo-2-{N-[2-(pyr­rol­idine-1-sul­fonyl)phenyl]car­box­imid­oyl}phenol, 20, em­erged with the smallest charge separation (ΔEgap = 3.5780 eV), signifying heightened potential for biological properties. Conversely, 4-bromo-2-{N-[2-(piper­idine-1-sul­fonyl)phenyl]car­box­imid­oyl}phenol, 17, exhibited the largest charge separation (ΔEgap = 4.9242 eV), implying a relatively lower propensity for biological activity. Moreover, the synthesized Schiff bases displayed re­marke­able inhibition of tankyrase poly(ADP-ribose) polymerase enzymes, integral in colon cancer, surpassing the efficacy of a standard drug used for the same purpose. Additionally, their bioavailability scores aligned closely with established medications such as trifluridine and 5-fluoro­uracil. The ex­plor­ation of mol­ecular electrostatic potential through colour mapping delved into the electronic behaviour and reactivity tendencies intrinsic to this diverse range of mol­ecules.

1. Introduction

The quest for effective anti­cancer agents remains a pivotal challenge in medicinal chemistry and pharmacology, particularly in the context of colon cancer, which is among the leading causes of cancer-related mortality worldwide (Kumar et al., 2023a[Kumar, A., Bhagat, K. K., Singh, A. K., Singh, H., Angre, T., Verma, A., Khalilullah, H., Jaremko, M., Emwas, A. H. & Kumar, P. (2023a). RSC Adv. 13, 6872-6908.],b[Kumar, A., Singh, A. K., Singh, H., Vijayan, V., Kumar, D., Naik, J. & Kumar, P. (2023b). Pharmaceuticals, 16, 299.]). The synthesis of novel com­pounds and the ex­plor­ation of their biological activities are critical steps in the development of new therapeutic agents. In this connection, aryl­sul­fon­amide Schiff bases have emerged as a class of com­pounds with significant potential due to their versatile chemical structures and promising pharmacological profiles (Irfan et al., 2020[Irfan, A., Rubab, L., Rehman, M. U., Anjum, R., Ullah, S., Marjana, M., Qadeer, S. & Sana, S. (2020). Heterocycl. Commun. 26, 46-59.]; Muhammad-Ali et al., 2023[Muhammad-Ali, M. A., Jasim, E. Q. & Al-Saadoon, A. H. (2023). J. Med. Chem. Sci. 6, 2128-2139.]; Dueke-Eze et al., 2020[Dueke-Eze, C. U., Fasina, T. M., Oluwalana, A. E., Familoni, O. B., Mphalele, J. M. & Onubuogu, C. (2020). Sci. Afr. 9, e00522.]).

Schiff bases, characterized by their imine functional group (–C=N–), have been studied extensively for their diverse pharmacological activities, including anti­cancer properties (Alblewi et al., 2023[Alblewi, F. F., Alsehli, M. H., Hritani, Z. M., Eskandrani, A., Alsaedi, W. H., Alawad, M. O., Elhenawy, A. A., Ahmed, H. Y., El-Gaby, M. S. A., Afifi, T. H. & Okasha, R. M. (2023). Int. J. Mol. Sci. 24, 16716.]). The introduction of an aryl­sul­fon­amide moiety into Schiff base structures has been hypothesized to enhance their biological activity, owing to the known efficacy of the sul­fon­amide group in various therapeutic agents (Abd El-Wahab et al., 2020[Abd El-Wahab, H., Abd El-Fattah, M., El-alfy, H. M. Z., Owda, M. E., Lin, L. & Hamdy, I. (2020). Prog. Org. Coat. 142, 105577.]; Elsamra et al., 2022[Elsamra, R. M., Masoud, M. S. & Ramadan, A. M. (2022). Sci. Rep. 12, 20192.]). The rationale behind this hypothesis centres on exploring the anti­cancer potential of Schiff bases, particularly focusing on their inter­action with colon cancer. Extensive literature has already underscored their efficacy as anti­cancer agents (Abd-Elzaher et al., 2016[Abd-Elzaher, M. M., Labib, A. A., Mousa, H. A., Moustafa, S. A., Ali, M. M. & El-Rashedy, A. A. (2016). Beni-Suef Univ. J. Basic Appl. Sci. 5, 85-96.]). Remarkably, Schiff bases have demonstrated activity against colon cancer and have been documented for such effects (Matela, 2020[Matela, G. (2020). Anticancer Agents Med. Chem. 20, 1908-1917.]). Notably, the combination of benzene­sul­fon­amide with a Schiff base has been reported, merging the bioactive attributes of sul­fon­amides and Schiff bases to investigate potential synergies between these well-established functional groups (Afsan et al., 2020[Afsan, Z., Roisnel, T., Tabassum, S. & Arjmand, F. (2020). Bioorg. Chem. 94, 103427.]). The cumulative evidence of their enhanced activity spurred our inter­est in undertaking the present study.

Tankyrase poly(ADP-ribose) polymerase, a crucial enzyme involved in DNA repair and the regulation of various cellular processes, has been implicated in the development of colon cancer (Feng & Koh, 2013[Feng, X. & Koh, D. W. (2013). Int. Rev. Cell Mol. Biol. 304, 227-281.]; Eisemann & Pascal, 2020[Eisemann, T. & Pascal, J. M. (2020). Cell. Mol. Life Sci. 77, 19-33.]). Tankyrase's involvement in Wingless-related integration site (Wnt) signaling, which governs cell growth, motility and differentiation, makes it a significant target (Pai et al., 2017[Pai, S. G., Carneiro, B. A., Mota, J. M., Costa, R., Leite, C. A., Barroso-Sousa, R. & Giles, F. J. (2017). J. Hematol. Oncol. 10, 101.]). Colorectal cancer, a prevalent form of cancer worldwide, often arises from precancerous polyps in the colon or rectum. Tankyrase's modification of Axin through poly(ADP-ribose) chains disrupts the Axin complex, leading to Axin degradation and β-catenin stabilization. The accumulation of β-catenin contributes to the progression of colon cancer (Gao et al., 2014[Gao, C., Xiao, G. & Hu, J. (2014). Cell Biosci. 4, 13.]). Under normal circumstances, Axin aids in regulating the Wnt pathway by facilitating β-catenin degradation (Huang & He, 2008[Huang, H. & He, X. (2008). Curr. Opin. Cell Biol. 20, 119-125.]). However, mutations in colon cancer can lead to the persistent accumulation of β-catenin, even in the absence of Wnt signaling, promoting uncontrolled cell growth and tumour formation (Behrens, 2000[Behrens, J. (2000). Ann. N. Y. Acad. Sci. 910, 21-35.]). Inhibiting specific amino acid residues in human tankyrase poly(ADP-ribose) polymerase using Schiff bases, such as (E)-2-[(2-hy­droxy­benzyl­idene)amino]­benzene­sul­fon­amide derivatives (1723) (Fig. 1[link]), could potentially prevent the accumulation of β-catenin, holding promise for effective inter­vention (Meyer et al., 2006[Meyer, R. G., Meyer-Ficca, M. L., Jacobson, E. L. & Jacobson, M. K. (2006). Enzymes in poly(ADP-ribose) metabolism, edited by A. Bürkle, pp. 1-12. Georgetown: Landes Bioscience.]).

[Figure 1]
Figure 1
General reaction scheme of the formation of potentially bioactive sul­fon­amide Schiff bases.

Our research also delves into the crystal structure of benzene­sul­fon­amides (Kolade et al., 2020[Kolade, S. O., Izunobi, J. U., Hosten, E. C., Olasupo, I. A., Ogunlaja, A. S. & Familoni, O. B. (2020). Acta Cryst. C76, 810-820.]), which, along with our ex­plor­ation of the crystal structure of a Schiff base sul­fon­amide, forms a comprehensive investigation into the potential therapeutic avenues these com­pounds may offer.

2. Experimental

2.1. Instruments and measurements

All reagents were purchased from Millipore Sigma (Ger­many and South Africa) and were used without further purification. The melting points were determined on an elec­tro­thermal digital melting-point apparatus and are uncorrected. Reactions were monitored by thin-layer chromatography (TLC) on Merck silica gel 60 F254 precoated plates using a di­chloro­methane/n-hexane (2 or 1.4:1 v/v) solvent system vis­ualized under a UV lamp (254 nm). Column chromatography was per­formed with silica gel (70–230 mesh ASTM) and mobile phases were as indicated. Sample crystallization was achieved by the slow evaporation of the indicated solvent systems at ambient temperature. IR spectra were obtained using a Bruker Tensor 27 platinum ATR–FT–IR spectrometer. The ATR–FT–IR spectra were acquired in a single mode with a resolution of 4 cm−1 over 32 scans in the region 4000–650 cm−1. 1H and 13C NMR spectra were record­ed in CDCl3 on a Bruker 400 MHz spectrometer. Chemical shift values were measured in parts per million (ppm) downfield from tetra­methyl­silane (TMS), and coupling constants (J) are reported in Hertz (Hz). Theoretical studies were per­formed for the com­pounds and, in each case, their single-crystal X-ray diffraction (SC-XRD) structures were used for optimization and global reactivity descriptor (GRD) calculations.

2.2. Synthesis and crystallization

2.2.1. Synthesis of sul­fon­amide Schiff bases

The general reaction scheme for the formation of potentially bioactive sul­fon­amide Schiff bases is shown in Fig. 1[link]. The N-cyclo­amino-2-sulf­anil­amides were prepared as reported previously (Kolade et al., 2022[Kolade, S. O., Izunobi, J. U., Gordon, A. T., Hosten, E. C., Olasupo, I. A., Ogunlaja, A. S., Asekun, O. T. & Familoni, O. B. (2022). Acta Cryst. C78, 730-742.]) by reacting the amino­sulf­anil­amides with substituted o-salicyl­aldehyde either at room temperature or under reflux to obtain the required Schiff bases in good yields (57–80%). Only o-salicyl­aldehyde and N-piperidinyl-2-sulf­anil­amide gave the required Schiff base at room temperature, and the others were refluxed to give the required products.

2.2.1.1. Method A: synthesis of N1-(21-hy­droxy­benzyl­iden­yl)-N-piperidinyl-2-sulf­anil­amide 17. N-Piperidinyl-2-sulf­anil­amide 11 (0.100 g, 0.417 mmol) was dissolved in methanol (5 ml) and 2-hy­droxy­benzaldehyde, or o-salicyl­aldehyde, 14 (0.056 g, 0.05 ml, 0.459 mmol), was added dropwise to the solution with stirring. Crushed ice (1.00 g) was added to the stirring mixture after 5 min. The reaction mixture was stirred at ambient temperature for 17 h and monitored with TLC. On completion, the mixture was filtered, using a Buckner funnel, and the residue was air-dried, dissolved in warm methanol and filtered hot to leave single crystals of 17 on slow evaporation. The physical properties and the spectroscopic data are pre­sent­ed in the supporting information.

2.2.1.2. Method B: synthesis of N1-(51-bromo/nitro-21-hy­droxy­benzyl­iden­yl)-N-cyclo­amino-2-sulf­anil­amides 1823. To a stirring solution of N-cyclo­amino-2-sulf­anil­amides 1013 (1.0 mmol) in ethanol (10 ml) was added 5-bromo­(nitro)-o-salicyl­aldehydes 1516 (1.3 mmol), followed by glacial acetic acid (10 drops) as catalyst. The whole mixture was refluxed for 48 h and monitored with TLC. After completion, the reaction mixture was allowed to cool to ambient temperature and kept in the fume hood for 24 h. The residue was then recrystallized from ethanol (10 ml) and filtered hot to leave crystals of 1823 on slow evaporation. The physical properties and the spectroscopic data are pre­sent­ed in the supporting information.

2.3. Docking studies

2.3.1. Selection of reference drugs and cancer protein macromolecule

Common anti­cancer standard drugs, such as capecitabine (ID: 60953), 5-fluoro­uracil (ID: 3385) and trifluridine (ID: 6256), were downloaded from Pubchem (https://pubchem.ncbi.nlm.nih.gov/, last accessed on May 25, 2023) and saved in .sdf format as reference to compare inhibitory per­formance with the synthesized chemical com­pounds. In order to evaluate the lead com­pounds as inhibitors of the tankyrase poly(ADP-ribose) polymerase family responsible for cancer pathogenesis (Shirai et al. 2020[Shirai, F., Mizutani, A., Yashiroda, Y., Tsumura, T., Kano, Y., Muramatsu, Y., Chikada, T., Yuki, H., Niwa, H., Sato, S., Washizuka, K., Koda, Y., Mazaki, Y., Jang, M. K., Yoshida, H., Nagamori, A., Okue, M., Watanabe, T., Kitamura, K., Shitara, E., Honma, T., Umehara, T., Shirouzu, M., Fukami, T., Seimiya, H., Yoshida, M. & Koyama, H. (2020). J. Med. Chem. 63, 4183-4204.]), its protein crystal structure (PDB entry 6kro) was downloaded from www.rcsb.org (last accessed on April 20, 2023).

2.3.2. Preparation of ligands, reference drugs and protein mol­ecules for docking

Synthesized com­pounds (drawn using Chemdraw 14.0 and saved in .sdf format) and the selected reference drugs saved as .sdf files were opened in PyRx 0.8 Autodock Vina software (Kondapuram et al., 2021[Kondapuram, S. K., Sarvagalla, S. & Coumar, M. S. (2021). In Molecular Docking for Comput-Aided Drug Design, pp. 463-477. New York: Academic Press.]). Energy minimization was car­ried out, followed by conversion into protein databank partial charge (pdbqt) ligands. The crystal structure of the protein mol­ecule tankyrase poly(ADP-ribose) polymerase at a resolution of 1.90 Å was also uploaded into BIOVIA Discovery Studio (Dassault Systémes, 2020[Dassault Systémes (2020). Discovery Studio. BIOVIA, San Diego, CA, USA. https://www.3ds.com/.]). The binding sites were determined and all unwanted heteroatoms and water mol­ecules were removed, while polar hydrogen bonds were added to give pure protein and saved as .pdb files (Pawar & Rohane, 2021[Pawar, S. S. & Rohane, S. H. (2021). Asian J. Res. Chem. 14, 86--88.]).

2.3.3. Mol­ecular docking

Docking simulations were per­formed with PyRx AutoDock using the Lamarkian genetic algorithm and default procedures for docking a flexible ligand to a rigid protein. Blind docking was initially per­formed to identify all potential binding sites on the target protein within a 90 × 75 × 75 cubic grid centre. A grid spacing of 1.00 Å was used for the calculation of the grid maps using the autogrid module of AutoDock tools. For each ligand, a set of nine independent runs were per­formed for the enzyme run against all ligands and reference drugs. Clear identification of the potential binding sites is followed by docking of ligands to the sites and the most probable and energetically favourable binding conformations were determined (Trott & Olson, 2010[Trott, O. & Olson, A. J. (2010). J. Comput. Chem. 31, 455-461.]). Docking solutions were analyzed and ranked on the basis of the Vina scoring functions. All calculations were car­ried out on PC-based machines running Microsoft Windows 10 operating systems. The resulting structures were vi­su­al­ized and analyzed using the Discovery Studio visualizer.

2.4. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 1[link]. Carbon-bound H atoms were added in idealized geometrical positions in a riding model. Nitro­gen-bound H atoms were located in a difference map and refined freely. The H atom of the hy­droxy group was allowed to rotate with a fixed angle around the C—O bond to best fit the experimental electron density. The Hirshfeld surface analyses were per­formed with CrystalExplorer (Version 21.5; 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.]).

Table 1
Experimental details for 18

Crystal data
Chemical formula C18H19BrN2O3S
Mr 423.32
Crystal system, space group Orthorhombic, Pbca
Temperature (K) 296
a, b, c (Å) 12.4070 (9), 17.4250 (14), 17.5276 (12)
V3) 3789.3 (5)
Z 8
Radiation type Mo Kα
μ (mm−1) 2.30
Crystal size (mm) 0.84 × 0.43 × 0.12
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Bruker, 2016[Bruker (2016). APEX2, SAINT, SADABS and TWINABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.133, 0.241
No. of measured, independent and observed [I > 2σ(I)] reflections 27860, 3356, 2255
Rint 0.060
(sin θ/λ)max−1) 0.597
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.062, 0.172, 1.09
No. of reflections 3356
No. of parameters 227
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.52, −0.38
Computer programs: APEX2 (Bruker, 2016[Bruker (2016). APEX2, SAINT, SADABS and TWINABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SAINT (Bruker, 2016[Bruker (2016). APEX2, SAINT, SADABS and TWINABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT2018 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2018 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), ShelXle (Hübschle et al., 2011[Hübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281-1284.]), ORTEP-3 for Windows (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]), PLATON (Spek, 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.]) and Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]).

3. Results and discussion

3.1. Chemistry

N1-(51-Substituted-21-hy­droxy­benzyl­iden­yl)-N-cyclo­amino-2-sulf­anil­amides 1723 were prepared from the reaction of N-cyclo­amino-2-sulf­anil­amides 1013 [as reported previously by Kolade et al. (2022[Kolade, S. O., Izunobi, J. U., Gordon, A. T., Hosten, E. C., Olasupo, I. A., Ogunlaja, A. S., Asekun, O. T. & Familoni, O. B. (2022). Acta Cryst. C78, 730-742.])] with substituted o-salicyl­aldehyde either at room temperature or under reflux to obtain the required Schiff bases in good yields (57–80%). Only o-sali­cyl­al­de­hyde and N-pi­peri­din­yl-2-sulf­anil­amide gave the re­quired Schiff base at room temperature, and the others were refluxed to obtain the desired products. The N1-(21-hy­droxy­benzyl­iden­yl)-N-piperidinyl-2-sulf­anil­amide 17, which was pre­pared at room temperature (Method A), was aimed at establishing an eco-friendly protocol. The reaction progress was monitored by TLC.

All the com­pounds synthesized were characterized by their melting points and their IR, 1H/13C NMR and MS spectra. In order to clarify the mode of bonding on the ligands, their IR spectra (as pre­sent­ed in the supporting information) confirm the formation of the sul­fon­amide Schiff base ligands 1723 by the appearance of a strong absorption band at around 1614–1618 cm−1, which is attributed to stretching vibrations of the azomethine group and the absence of the original aldehydic bond (–C=O) and NH2 vibrations (Salehi et al., 2019[Salehi, M., Kubicki, M., Galini, M., Jafari, M. & Malekshah, R. E. (2019). J. Mol. Struct. 1180, 595-602.]). The stretching vibrations of aromatic carbon-to-carbon double bonds (C=C) of the com­pounds are observed at 1512–1570 cm−1, while the strong absorption bands which appeared at around 1300–1331 and 1119–1155 cm−1 are indexed to (S=O)2 asymmetric and symmetric stretching frequencies, respectively. The IR spectra provided in the supporting information also reveal other diagnostic bands that further corroborate the formation of Schiff base ligands.

The 1H NMR spectra (see supporting information) of sul­fon­amide Schiff bases (ligands) 1723 were record­ed in CDCl3, using tetra­methyl­silane (TMS) as the inter­nal standard. The signals at 1.36–4.46 ppm in the 1H NMR spectra of the ligands result from the protons of the methyl­ene groups (–CH2–). The singlet signals which correspond to the imine groups (–CH=N–) in these ligands are observed at 8.12–8.67 ppm. The phenolic protons (–OH), the most deshielded protons, are clearly indicated at 12.31–13.50 ppm. The deshielded nature of the phenolic OH hydrogen is likely a consequence of it forming a strong resonance-assisted intramolecular O—H⋯N hydrogen bond. The aromatic protons of the com­pounds are record­ed in the range 6.66–8.44 ppm. Finally, the success of the formation of the sul­fon­amide Schiff bases is corroborated by the 13C NMR spectra (see supporting information) of com­pounds 1723, which show the azomethine C atoms (–C=N–) at the chemical environments of 160–162 ppm, while the most deshielded phenolic C atoms occur at 163–166 ppm and the aromatic C atoms are observed at 111–148 ppm.

3.2. Crystal structure

Compound 18 formed pale-yellow platelets with the ortho­rhom­bic space group Pbca (Table 1[link]). The close ortho positioning of the two functional groups on the central arene ring forces their rotation, with a resulting dihedral angle of 30.8 (2)° for the least-squares planes through the piperidine and the imino­methyl­phenol groups, and dihedral angles of 77.17 (17) and 51.82 (12)°, respectively, with the central linking arene group. The imino­methyl­phenol group is planar, with an intra­molecular O—H⋯N inter­action of 1.86 Å, forming a ring closure that can be described with an S(6) graph-set descriptor (Table 2[link]). The inter­molecular hydrogen-bond inter­actions are dominated by the O atoms from the sulfonyl group, with a number of C—H⋯O=S inter­actions, resulting in three chains of inter­actions having C(7), C(9) and C(7) descriptors. The Hirshfeld surface illustrated in Fig. S1 (see supporting information) clearly shows these inter­actions. The hy­droxy group also contributes and is involved in a C—H⋯O inter­action of 2.54 Å, resulting in a C(8) chain inter­action. The presence of these C—H⋯O inter­actions is indicated on the Hirshfeld surface fingerprint plot as H⋯O (see Fig. S2). There is also an inter­molecular C—H⋯π(ring) inter­action of 2.88 Å to the centroid of the C11–C16 ring. This inter­action is indicated by H⋯C on the Hirshfeld surface fingerprint plot (Fig. S2). Table S3 (see supporting information) lists the percentage reciprocal hydrogen surface contact areas, with the H⋯H inter­actions having the largest percentage contact. The closest H⋯H contact indicated in Fig. S2 arises between H atoms on the arene rings between two ad­ja­cent imino­methyl­phenol groups. The H⋯O/O⋯H and H⋯C/C⋯H inter­actions have similar contact surface areas, while H⋯Br/Br⋯H inter­actions are also present.

Table 2
Hydrogen-bond geometry for 18 (Å, °)

Cg is the centroid of the C11–C16 ring.

D—H⋯A D—H H⋯A DA D—H⋯A
O1—H1A⋯N1 0.82 1.86 2.586 (5) 146
C1—H1⋯O2i 0.93 2.54 3.363 (6) 149
C16—H16⋯O2i 0.93 2.64 3.461 (6) 147
C23—H23⋯O3 0.93 2.43 2.846 (7) 107
C25—H25⋯O3ii 0.93 2.49 3.220 (6) 135
C26—H26⋯O1i 0.93 2.62 3.467 (7) 151
C35—H35B⋯O2 0.97 2.43 2.852 (7) 106
C32—H32BCgiii 0.97 2.88 3.664 (7) 139
Symmetry codes: (i) [x+{\script{1\over 2}}, y, -z+{\script{1\over 2}}]; (ii) [x+{\script{1\over 2}}, -y+{\script{1\over 2}}, -z]; (iii) [-x+{\script{3\over 2}}, y-{\script{1\over 2}}, z].

3.3. Theoretical calculations

3.3.1. DFT calculations

Mol­ecular orbital calculations, en­com­pas­sing full geometry optimization, were methodically conducted on the Schiff base derivatives alongside the established pharmaceutical com­pounds trifluridine, capecitabine and 5-fluoro­uracil. These sophisticated calculations were executed using the Jaguar module within Maestro (Version 13.6.122) and MMshare (Version 6.2.122, Release 2023-2). This involved the integration of the basis set 6-31G* level, harmonized with the hybrid density functional theory (DFT) that incorporates the Becke 3-parameter exchange potential (Becke, 1993[Becke, A. D. (1993). J. Chem. Phys. 98, 5648-5652.]; Prokopenko et al., 2019[Prokopenko, Y. S., Perekhoda, L. O. & Georgiyants, V. A. (2019). J. Appl. Pharm. Sci. 9, 066-072.]; Jędrzejczyk et al., 2022[Jędrzejczyk, M., Janczak, J. & Huczyński, A. (2022). J. Mol. Struct. 1263, 133129.]; Pandi et al., 2022[Pandi, S., Kulanthaivel, L., Subbaraj, G. K., Rajaram, S. & Subramanian, S. (2022). BioMed Res. Int. 2022, 3338549.]). This intricate approach paved the way for the meticulous determination of crucial mol­ecular properties. The focus of this investigation involved the precise computation of the highest occupied mol­ecular orbital (HOMO) and the lowest unoccupied mol­ecular orbital (LUMO) using the aforementioned methodology. The outcomes of these meticulous calculations, which illuminate the intricate electronic structure and reactivity of the mol­ecules, served as the foundation for the sub­se­quent computation of pivotal global reactivity descriptors. These encompass a spectrum of descriptors, notably the ionization po­ten­tial (I), electron affinity (A), chemical po­ten­tial (μ), electronegativity (χ), global hardness (η), global softness (S) and global electrophilicity (ω) values (Gordon et al., 2022[Gordon, A. T., Hosten, E. C. & Ogunlaja, A. S. (2022). Pharmaceuticals, 15, 1240.]).

3.3.2. Global reactivity descriptors of the synthesized Schiff bases and standard drugs

Density functional theory (DFT) stands as a widely embraced technique for ab-initio assessments of diverse mol­ecular components. Among its manifold utilities, it holds prominence in discerning the characteristics of frontier mol­ecular orbitals (FMOs), a pivotal factor in elucidating various reaction types and predicting the most reactive sites within conjugated systems (da Silva et al., 2006[Silva, R. R. da, Ramalho, T. C., Santos, J. M. & Figueroa-Villar, J. D. (2006). J. Phys. Chem. A, 110, 1031-1040.]). This comprehension of structure–property relationships assumes paramount importance in the endeavour to craft enhanced pharmaceutical agents, given that the mol­ecular configuration profoundly influences the per­formance of drugs (Mahmood et al., 2022[Mahmood, A., Irfan, A. & Wang, J. L. (2022). Chem. Eur. J. 28, e202103712.]).

The bedrock for global vital reactivity descriptors lies in the FMO properties, precisely, the HOMO and LUMO energy values. Through a judicious application of DFT energetics, this study delved into the intricate tapestry of three-dimensional electronic states intrinsic to the mol­ecules under scrutiny. As such, the analysis offered an unprecedented glimpse into the transferability of lone pairs, the nuances of bond inter­actions, and the reactivity landscape within the specific mol­ecular milieu (Hall et al., 2009[Hall, M. L., Goldfeld, D. A., Bochevarov, A. D. & Friesner, R. A. (2009). J. Chem. Theory Comput. 5, 2996-3009.]). In its totality, this exhaustive computational inquiry provides an illuminating vista into the intricate electronic characteristics and reactivity proclivities of the mol­ecules under examination. The insights gleaned from this study significantly enrich our understanding of their potential roles and behaviours across a spectrum of chemical scenarios.

In the present context, the com­pounds under scrutiny underwent a meticulous ex­plor­ation of their quantum chemi­cal attributes. Specifically, the focus was on the localization energies of the HOMO and the LUMO. These energies, encapsulated within the rubric of FMOs, serve as linchpins in upholding chemical stability. Moreover, they emerge as potent tools for dissecting donor–acceptor inter­actions. The HOMO embodies a mol­ecule's capacity to donate an electron, while the LUMO signifies its propensity to accept an electron. A lower LUMO value indicates an augmented inclination for electron acceptance, while higher HOMO values delineate a heightened disposition to donate electrons to unoccupied mol­ecular orbitals (Yele et al., 2021[Yele, V., Sigalapalli, D. K., Jupudi, S. & Mohammed, A. A. (2021). J. Mol. Model. 27, 359.]). Considering the context of the LUMO within the synthesized Schiff bases, 19 displays the most intriguing attribute, featuring the lowest energy level for the LUMO orbital (−2.4963 eV), indicative of its pronounced tendency to accept electrons. Conversely, 17 showcases the highest LUMO energy level (−1.3287 eV). This establishes an order of increasing electron-accepting tendency among the com­pounds: 19 > 20 > 23 > 21 > 18 > 22 > 23 > capecitabine > trifluridine > 17 > 5-fluoro­uracil (Table 3[link]).

Table 3
Global reactivity descriptors of the synthesized Schiff bases and standard drugs

Entry EHOMO ELUMO ΔEgap I A μ X η S ω
17 −6.2529 −1.3287 4.9242 6.2529 1.3287 −3.7908 3.7908 2.4621 0.4062 2.9183
20 −5.9851 −2.4071 3.5780 5.9851 2.4071 −4.1961 4.1961 1.7890 0.5590 4.9210
21 −6.1535 −2.0050 4.1485 6.1535 2.0050 −4.0793 4.0793 2.0743 0.4821 4.0112
18 −6.1546 −2.0050 4.1496 6.1546 2.0050 −4.0798 4.0798 2.0748 0.4820 4.0112
19 −6.5745 −2.4963 4.0782 6.5745 2.4963 −4.5354 4.5354 2.0391 0.4904 5.0439
22 −6.1778 −1.6035 4.5743 6.1778 1.6035 −3.8907 3.8907 2.2872 0.4372 3.3092
23 −6.6167 −2.2713 4.3454 6.6167 2.2713 −4.4440 4.4440 2.1727 0.4603 4.5448
5-Flu −6.7827 −1.2659 5.5168 6.7827 1.2659 −4.0243 4.0243 2.7584 0.3625 2.9355
Cap −6.4578 −1.6019 4.8559 6.4578 1.6019 −4.0299 4.0299 2.4279 0.4119 3.3444
Tri −6.9868 −1.3725 5.6143 6.9868 1.3725 −4.1797 4.1797 2.8071 0.3562 3.1117
Notes: ΔEgap is the energy gap or charge separation, I is ionization potential, A is electron affinity, μ is chemical potential, X is electronegativity, η is global hardness, S is global softness and ω is the electrophilicity index. Tri is trifluridine, Cap is capecitabine and 5-Flu is 5-fluoro­uracil.

On a contrasting note, when focusing on EHOMO, 20 de­mon­strates the highest energy level (−5.9851 eV), followed closely by 22 with the second highest EHOMO (−6.1778 eV). These values signify the pronounced potential of 20 and 22 to donate electrons (Yele et al., 2021[Yele, V., Sigalapalli, D. K., Jupudi, S. & Mohammed, A. A. (2021). J. Mol. Model. 27, 359.]). Remarkably, capecitabine emerges as the only standard drug displaying HOMO energies higher than the synthesized com­pounds (19 and 23), boasting an energy level of EHOMO = −6.4578 eV. Beyond mere energy levels, a thorough structural analysis encompasses a comprehensive evaluation of intra-ligand inter­actions. Notably, these inter­actions include hydrogen bonds (within a 3.5 Å range), halogen bonds (within a 3.5 Å range), ππ stacking (within a 5.5 Å range) and π–cation inter­actions (within a 6.6 Å range).

Delving deeper into the mol­ecular architecture, both the HOMO and the LUMO orbitals exhibit localization on both arene rings of 17. However, no discernible hydrogen bonds or ππ stacking within the studied distances are exhibited by the com­pound. In the intricate case of 20, HOMO orbitals predominantly localize on the arene ring bearing the Br atom, while the LUMO orbitals are positioned closer to the imine functionality (C=N). This arrangement leads to the identification of hydrogen bonding within the optimized structure of 20, specifically involving C=N⋯OH (1.50 Å) and a weaker S=O⋯H(aromatic) inter­action (Fig. 2[link]).

[Figure 2]
Figure 2
Frontier mol­ecular orbitals (FMOs) of 17 and 20.

Similarly, Schiff base 21 (Fig. 3[link]) showcases HOMO orbitals predominantly on the bromine-bearing arene ring, while the LUMO orbitals position themselves in proximity to the imine functionality. Notably, 21 boasts robust hydrogen bonding between the hy­droxy O and imine N atom (1.50 Å). Extending this pattern, derivatives 18 and 19 exhibit comparable hydrogen bonding to 21, with distances of 1.70 and 1.73 Å, respectively. Schiff base 19 (Fig. 4[link]) distinguishes itself further by showcasing an additional, albeit weaker, hydrogen bond (2.67 Å) between the S=O group and an aromatic H atom. The thematic consistency in the HOMO and LUMO orbital localization is mirrored across Schiff bases 18, 19, 22 and 23, closely resembling the pattern exhibited by 20, except for 23, where the LUMO orbitals predominantly localize on the NO2 group (Fig. 5[link]).

[Figure 3]
Figure 3
Frontier mol­ecular orbitals (FMOs) of 21 and 18.
[Figure 4]
Figure 4
Frontier mol­ecular orbitals (FMOs) of 19 and 22.
[Figure 5]
Figure 5
Frontier mol­ecular orbitals (FMOs) of 23 and capecitabine.

Within the structures of 5-fluoro­uracil and trifluridine (Fig. 6[link]), the HOMO and LUMO exhibit localization in distinct regions of each respective mol­ecule. This localization directly signifies the occurrence of charge-transfer processes.

[Figure 6]
Figure 6
Frontier mol­ecular orbitals (FMOs) of 5-fluoro­uracil and trifluridine.

Turning our focus to broader implications, the eigenvalues of the HOMO and LUMO, along with their energy gap, offer crucial insights into the biological activity of a molecule (Table 3[link]). A diminished energy gap, symbolized as ΔEgap, renders a mol­ecule more susceptible to polarization. This phenomenon aligns with heightened chemical reactivity and reduced kinetic stability, ultimately driving positive impetus toward biological activity. In contrast, an enlarged energy gap between the HOMO and LUMO orbitals signifies the kinetic instability of a mol­ecule, translating to a diminished propensity for biological activity (Pereira et al., 2017[Pereira, F., Xiao, K., Latino, D. A., Wu, C., Zhang, Q. & Aires-de-Sousa, J. (2017). J. Chem. Inf. Model. 57, 11-21.]; Akman, 2019[Akman, F. (2019). Cellul. Chem. Technol. 53, 243-250.]; Choudhary et al., 2013[Choudhary, N., Bee, S., Gupta, A. & Tandon, P. (2013). Comput. Theor. Chem. 1016, 8-21.]; Abdelsalam et al., 2022[Abdelsalam, M. M., Bedair, M. A., Hassan, A. M., Heakal, B. H., Younis, A., Elbialy, Z. I., Badawy, M. A., Fawzy, H. E. & Fareed, S. A. (2022). Arabian J. Chem. 15, 103491.]). Adding a layer of nuance, 20 emerges as the mol­ecule showcasing the smallest charge separation (ΔEgap = 3.5780 eV), suggesting its heightened potential for biological properties. In contrast, 17 displays the largest charge separation (ΔEgap = 4.9242 eV), indicating a comparatively lower propensity for biological properties compared to the other synthesized Schiff bases.

In essence, the meticulous unravelling of these quantum features through DFT provides a profound understanding of the intricate inter­play between mol­ecular structure, reactivity and biological per­formance. Such insights hold transformative potential in advancing drug design and precision chemical manipulation. The distinctiveness of ligand 20 is further underscored by its characterization as the `softest' mol­ecule (S = 0.5589 eV) and, consequently, the `least hard' mol­ecule (η = 1.7890 eV). Conversely, 17 exhibits the highest hardness value (η = 2.4621 eV) and lowest softness (S = 0.4062 eV). When we turn our attention to standard drugs, the order of softness is observed as capecitabine > 5-fluoro­uracil > trifluridine, with all values generally lower than most synthesized Schiff bases. The chemical potential (μ) spans from −4.5354 eV (lowest) for 19 to −3.7908 eV (highest) for 17. Schiff bases 19 and 20 exhibit the highest electrophilicity index (ω), with values of 5.044 and 4.9209 eV, respectively. In contrast, Schiff base 17 showcases the lowest electrophilicity index (ω = 2.9182 eV).

With regard to the computed global reactivity indices and the HOMO–LUMO gap, the order is: 20 > 19 > 21 > 18 > 23 > 22 > capecitabine > 17 > 5-fluoro­uracil > trifluridine. This insightful hierarchy provides valuable direction for the reactivity and potential biological activity of the synthesized Schiff bases.

3.3.3. Mol­ecular Electrostatic Potential (MESP)

The Mol­ecular Electrostatic Potential (MESP) concept serves as a window into the intricate charge distribution enveloping mol­ecules within the expanse of three-dimensional space. Its significance is particularly pronounced in identifying susceptible loci for electrophilic and nucleophilic inter­actions, which are critical in the realm of biological recognition and hydrogen-bonding phenomena. Through the utilization of colour mapping grounded in electron density, the electrostatic potential of the studied mol­ecules found visual expression, as illustrated in Figs. 2–6[link][link][link][link][link].

This visual representation employs a spectrum of colours to delineate the MESP surface characteristics. Red hues signify regions enriched with electrons, indicating a partially negative charge, while blue shades indicate electron-deficient zones with a partial positive charge. Light-blue nuances mark slightly electron-deficient areas, while yellow tinges highlight slightly electron-rich regions. Neutral zones with a zero potential are depicted in green (Altürk et al., 2015[Altürk, S., Tamer, Ö., Avcı, D. & Atalay, Y. (2015). J. Organomet. Chem. 797, 110-119.]; Friesner et al., 2006[Friesner, R. A., Murphy, R. B., Repasky, M. P., Frye, L. L., Greenwood, J. R., Halgren, T. A., Sanschagrin, P. C. & Mainz, D. T. (2006). J. Med. Chem. 49, 6177-6196.]).

Upon scrutinizing the MESP mappings of individual com­pounds, distinct patterns emerge. Schiff base 17 [Fig. 7[link](a)] pre­dominantly reveals a green surface, save for the hy­droxy-functionalized section which distinctly appears in blue. Similarly, the MESP profile of com­pound 20 [Fig. 7[link](b)] features prevalent blue regions, with a specific green–yellow region localized over the bromine-bearing arene ring. In the case of 21 [Fig. 8[link](a)], an evident gradient from green to blue characterizes the MESP map. Analogously, the MESP portrayal of 18 [Fig. 9[link](b)] reflects this trend, except for the S=O functional-group regions which assume a red hue. Both 19 [Fig. 9[link](a)] and 22 [Fig. 9[link](b)] display a blending of blue and green regions in their respective MESP renderings. For 23 [Fig. 10[link](a)], the MESP map predominantly features green hues, while the hy­droxy-enriched area adopts a distinctive blue shade. Noteworthy instances include the standard drug capecitabine [Fig. 10[link](b)], predominantly depicted in blue in its MESP representation. The standard drug 5-fluoro­uracil [Fig. 11[link](a)] showcases the entire spectrum of colour variations across its surface. Similarly, trifluridine [Fig. 11[link](b)] transitions from blue to red, thereby illustrating its surface characteristics encompassing the 2-(hy­droxy­meth­yl)tetra­hy­dro­furan-3-ol moi­ety.

[Figure 7]
Figure 7
MESP plots of (a) 17 and (b) 20. Regions of attractive potential appear in red and those of repulsive potential appear in blue.
[Figure 8]
Figure 8
MESP plots of (a) 21 and (b) 18.
[Figure 9]
Figure 9
MESP plots of (a) 19 and (b) 22.
[Figure 10]
Figure 10
MESP plots of (a) 23 and (b) capecitabine.
[Figure 11]
Figure 11
MESP plots of (a) 5-fluoro­uracil and (b) trifluridine.

3.4. Docking studies

3.4.1. Schiff bases as potential inhibitors of tankyrase colon cancer protein mol­ecules

The docking car­ried out using AutoDock Vina on the PyRx website (https://pyrx.sourceforge.io/) and the summary of the binding energy of each ligand obtained is pre­sent­ed in Table 4[link] (the structures are pre­sent­ed in Fig. 12[link]). In addition, the drug-likeness and toxicity of the synthesized Schiff bases and reference drugs were also investigated (Tables 5–10[link][link][link][link][link][link]).

Table 4
Summary of the binding energy (kcal mol−1) of Schiff bases with poly(ADP-ribose) polymerase

Optimized Schiff bases Summarized drug-likeness and toxicity 6kro
6kro_23_E=714.59 mildly nondrug-like and toxic −11.1
6kro_22_E=797.81 nontoxic but mildly nondrug-like −10.3
6kro_21_E=687.32 nondrug-like and nontoxic −9.9
6kro_20_E=635.91 drug-like and nontoxic −9.5
6kro_17_E=666.05 drug-like and nontoxic −9.2
6kro_18_E=685.47 drug-like and nontoxic −8.7
6kro_19_E=748.77 drug-like but mildly toxic −6.7
6kro_trifluridine_E=282.80 drug-like but mildly toxic −8
6kro_capecitabine_E=624.15 drug-like but highly toxic −7.9
6kro_5-fluoro­uracil_E=45.84 mildly nondrug-like and toxic −5.5

Table 5
In-silico toxicity study and drug-likeness of 20 using ProTox-II and Swiss­ADME

Druglikeness
[Scheme 3]
Toxicity
Target Prediction Probability
Hepatotoxicity Inactive 0.57
Carcinogenicity Inactive 0.57
Immunotoxicity Inactive 0.87
Mutagenicity Inactive 0.71
Cytotoxicity Inactive 0.73

Table 6
In-silico toxicity study and drug-likeness of 23 using ProTox-II and Swiss­ADME

Druglikeness
[Scheme 2]
Toxicity
Target Prediction Probability
Hepatotoxicity Inactive 0.62
Carcinogenicity Inactive 0.57
Immunotoxicity Inactive 0.94
Mutagenicity Active 0.68
Cytotoxicity Inactive 0.78

Table 7
In-silico toxicity study and drug-likeness of trifluridine

Druglikeness
[Scheme 1]
Toxicity
Target Prediction Probability
Hepatotoxicity Inactive 0.76
Carcinogenicity Inactive 0.60
Immunotoxicity Inactive 0.99
Mutagenicity Active 0.64
Cytotoxicity Inactive 0.88

Table 8
In-silico toxicity study and drug-likeness of 5-fluoro­uracil

Druglikeness
[Scheme 4]
Toxicity
Target Prediction Probability
Hepatotoxicity Inactive 0.78
Carcinogenicity Active 0.85
Immunotoxicity Inactive 0.99
Mutagenicity Inactive 0.88
Cytotoxicity Inactive 0.93

Table 9
Physicochemical properties of the synthesized Schiff bases and reference drugs

Compound Mr No. of heavy atoms Fraction Csp3 No. rotational bonds No. hydrogen-bond acceptors No. hydrogen-bond donors TPSA log Kp (cm s−1) Bioavailability score
17 344.43 24 0.28 4 5 1 78.35 −6.37 0.55
20 409.3 24 0.24 4 5 1 78.35 −6.53 0.55
21 457.34 28 0.10 4 4 1 78.35 −6.04 0.55
18 423.32 25 0.28 4 5 1 78.35 −6.36 0.55
19 389.43 27 0.28 5 7 1 124.17 −6.77 0.55
22 471.37 29 0.14 4 5 1 78.35 −6.17 0.55
23 437.47 31 0.14 5 7 1 124.17 −6.58 0.55
Trifluridine 296.2 20 0.60 3 8 3 104.55 −8.43 0.55
Capecitabine 359.35 25 0.67 8 8 3 122.91 −8.09 0.55
5-Fluoro­uracil 130.08 9 0 0 3 2 65.72 −7.73 0.55

Table 10
Drug-likeness rule violation and in-silico toxicity study of trifluridine and 5-fluoro­uracil for comparison

GI is gastrointestinal and BBB is blood–brain barrier.

Drug-likeness rules violations Blood–brain distribution and metabolism
Compounds Lipinski Ghose Veber Egan Muegge GI absorption BBB permeant
17 0 0 0 0 0 High No
20 0 0 0 0 0 High No
21 0 0 0 0 0 High No
18 0 0 0 0 0 High No
19 0 0 0 0 0 High No
22 0 0 0 0 0 Low No
23 0 0 0 0 0 High No
Trifluridine 0 0 0 0 0 High No
Capecitabine 0 0 0 0 0 High No
5-Fluoro­uracil 0 0 0 0 0 High No
[Figure 12]
Figure 12
Structures of the synthesized Schiff bases 1723 used for docking.

It is noteworthy that ligand 23 has the highest binding energy of −11.1 kcal mol−1, probably due to the presence of not only the sul­fon­amide but also the nitro group coupled with the tetra­hydro­iso­quinoline moiety binding to the protein.

The binding energy was observed to be significantly higher than that of the reference drugs trifluridine, capecitabine and 5-fluoro­uracil having binding energies of −8.0, −7.9 and −5.5 kcal mol−1, respectively. However, when the toxicity and the drug-likeness of ligand 23 were checked using the ProTox-II webserver and Swiss­ADME (http://www.swissadme.ch/), respectively, it was toxic and failed some of the rules despite its excellent binding inter­action.

Inter­estingly, ligands 17, 20 and 18 were completely non­toxic and followed all drug-likeness rules, unlike all the other synthesized Schiff bases (i.e. 21, 19, 22 and 23; see supporting information). In comparison, the reference drugs were also toxic, falling short of at least one drug-likeness rule. The best reference drug, i.e. trifluridine, and ligand 20, namely, (E)-4-bromo-2-({[2-(pyrrolidin-1-ylsulfon­yl)phen­yl]imino}­meth­yl)phenol, with the best binding energy and in compliance with all drug-like rules and displaying complete nontoxicity, were selected for further study (Table 10[link]).

The 2D and 3D structures of 20 showing the inter­acting amino acid residues [Fig. 13[link](a)], bond lengths [Fig. 13[link](b)], hydro­phobic inter­actions [Fig. 13[link](c)] and solvent-accessibility surface [Fig. 13[link](d)] are all pre­sent­ed. One of the sul­fon­amide O atoms exhibits a hydrogen-bonding inter­action with amino acid residue Arg1100 at a bonding distance of 1.97 Å, which is also noticeable within the atoms of the ligand in an intra­molecular fashion [Figs. 13[link](a) and 13(b)]. π-Alkyl and T-shaped inter­actions were also exhibited between the pyrrolidine moiety and amino acid residues Val1000 and Leu1097, and between the π-electrons of the two aromatic rings and the Trp1006 and Tyr1009 residues, respectively.

[Figure 13]
Figure 13
2D and 3D structures of synthesized Schiff base 20 showing (a) the inter­acting amino acid residues, (b) bond lengths, (c) hydro­phobic inter­actions and (d) solvent-accessibility surface.

Significantly, the amino acid residues inter­acting with tankyrase poly(ADP-ribose) polymerase residues prefer hydro­phobic inter­actions [as depicted by the deep-brown region of Fig. 13[link](c)]. While Arg1100 had good solvent-accessibility surface inter­actions, other inter­acting residues had excellent solvent inter­actions with ligand 20 [Figs. 13[link](c) and 13(d)]. Although the reference drug trifluridine has two hydrogen-bond inter­actions, they are comparatively weaker and have longer bond lengths of 2.14 and 2.60 Å with Gly1032 and Asp1045, respectively, when compared with ligand 20 (1.90 Å), as pre­sent­ed in Figs. 14[link](a) and 14(b).

[Figure 14]
Figure 14
2D and 3D structures of the reference drug trifluridine, showing (a) the inter­acting amino acid residues, (b) bond lengths, (c) hydro­phobic inter­actions and (d) solvent-accessibility surface.

Noticeably, the solvent-accessibility surface of the reference drug seems better, as all inter­acting amino acid residues inter­act in the blue region [Fig. 14[link](d)]; however, it has comparably lower binding energy (Table 4[link]), i.e. poorer hydro­phobicity than exhibited by most drug-like candidates [Fig. 14[link](c)], and it possesses mutagenic toxicity (Table 7[link]).

3.4.2. Toxicity and drug-likeness of Schiff bases 17–23 and reference drugs

The SMILES (simplified mol­ecular-input line-entry system) of the synthesized Schiff bases and the reference drugs were obtained via ChemDraw (Version 14.0) software and PubChem, respectively. These SMILES were uploaded into the online webservers Pro-Tox-II and Swiss­ADME to investigate the in-silico toxicity and drug-likeness. A summary of the results obtained is pre­sent­ed in Tables 5–10[link][link][link][link][link][link]. From the results, it is obvious that ligand 23 (−11.10 kJ mol−1) fell short of the toxicity test, despite being the best inter­acting ligand (Table 7[link]). It could also not completely fit into the hexa­gonal drug-likeness physicochemical space. From the investigation, it became clear that ligand 20, with a binding energy of −9.50 kJ mol−1, is completely nontoxic and fits perfectly into the hexa­gon, thereby displaying 100% drug-likeness (Table 6[link]).

In comparison, the two common colon cancer reference drugs used in this study show some levels of toxicity. While trifluridine is mildly mutagenic, 5-fluoro­uracil is highly carcinogenic (Tables 8[link] and 9[link]). Unlike synthesized Schiff bases 17, 20 and 18, this study also reveals that 5-fluoro­uracil fails some drug-likeness tests in addition to its toxic nature (Table 9[link]).

In the course of the drug-likeness investigation, physicochemical parameters and drug-likeness violations of the Schiff bases and the reference drugs trifluridine and 5-fluoro­uracil were also compared, as shown in Tables 9[link] and 10[link]. Most of the properties, such as the number of heavy atoms, rotatable bonds, TPSA (topological polar surface area), log Kp and bioavailability scores of the synthesized ligands compare effectively with trifluridine and 5-fluoro­uracil. Inter­estingly, none of the synthesized ligands violated Absorption, Distribution, Metabolism and Excretion (ADME) rules; hence, they can be tagged as potential drug candidates. While their gastrointestinal (GI) absorption is very high, the same properties exhibited by the reference drugs, none of the Schiff bases are blood–brain barrier permeant, making them safe without any unwarranted inter­ference with the central nervous system (Table 9[link]).

4. Conclusion

The successful synthesis, characterization and analysis of the inter­molecular inter­actions of N1-(51-substituted-21-hy­droxy­benzyl­iden­yl)-N-cyclo­amino-2-sulf­anil­amides (com­pounds 1723) have been achieved, alongside their evaluation for inhibitory effects on tankyrase poly(ADP-ribose) polymerase in the context of colon cancer through in-silico testing. Crystal packing and density functional theory (DFT) analyses have indicated that hydrogen bonds and ππ stacking play a crucial role in the mol­ecular cohesion of these com­pounds. Furthermore, the DFT results, when combined with mol­ecular docking studies, reveal that the electronegativity and electrophilicity attributes of these com­pounds significantly influence their binding affinity towards tankyrase poly(ADP-ribose) polymerase. This comprehensive study not only sheds light on the underlying mechanisms of action but also lays down a foundational framework for the development of effective therapies against colon cancer based on com­pounds 1723.

Supporting information


Computing details top

4-Bromo-2-{N-[2-(piperidine-1-sulfonyl)phenyl]carboximidoyl}phenol top
Crystal data top
C18H19BrN2O3SDx = 1.484 Mg m3
Mr = 423.32Melting point: 427.45 K
Orthorhombic, PbcaMo Kα radiation, λ = 0.71073 Å
a = 12.4070 (9) ÅCell parameters from 9105 reflections
b = 17.4250 (14) Åθ = 2.3–24.5°
c = 17.5276 (12) ŵ = 2.30 mm1
V = 3789.3 (5) Å3T = 296 K
Z = 8Platelet, yellow
F(000) = 17280.84 × 0.43 × 0.12 mm
Data collection top
Bruker APEXII CCD
diffractometer
3356 independent reflections
Radiation source: sealed tube2255 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.060
Detector resolution: 8.3333 pixels mm-1θmax = 25.1°, θmin = 2.3°
φ and ω scansh = 1414
Absorption correction: multi-scan
(SADABS; Bruker, 2016)
k = 1920
Tmin = 0.133, Tmax = 0.241l = 2020
27860 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)] = 0.062Hydrogen site location: inferred from neighbouring sites
wR(F2) = 0.172H-atom parameters constrained
S = 1.09 w = 1/[σ2(Fo2) + (0.061P)2 + 7.1303P]
where P = (Fo2 + 2Fc2)/3
3356 reflections(Δ/σ)max < 0.001
227 parametersΔρmax = 0.52 e Å3
0 restraintsΔρmin = 0.37 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.

Refinement. Carbon-bound H atoms were placed in calculated positions and were included in the refinement in the riding model approximation, with Uiso(H) set to 1.2 Ueq(C).

The H atom of the hydroxyl group was allowed to rotate with a fixed angle around the C—O bond to best fit the experimental electron density (HFIX 147 in the SHELXL program (Sheldrick, 2015)), with Uiso(H) set to 1.5Ueq(O).

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.85040 (8)0.46864 (5)0.55005 (4)0.1140 (4)
S10.56327 (10)0.22940 (8)0.14642 (7)0.0607 (4)
O10.5496 (3)0.4349 (3)0.2834 (3)0.0907 (13)
H1A0.5795490.4135410.2475490.136*
O20.4882 (3)0.2902 (2)0.1591 (2)0.0777 (11)
O30.5316 (3)0.1682 (2)0.0968 (2)0.0883 (12)
N10.6996 (3)0.3585 (2)0.2148 (2)0.0555 (10)
N20.5983 (3)0.1974 (2)0.2293 (2)0.0565 (10)
C10.7606 (4)0.3689 (3)0.2721 (3)0.0533 (11)
H10.8309120.3504730.2706880.064*
C110.7232 (4)0.4088 (3)0.3396 (3)0.0543 (12)
C120.6188 (5)0.4412 (3)0.3421 (3)0.0688 (15)
C130.5885 (5)0.4815 (3)0.4058 (4)0.091 (2)
H130.5206180.5040490.4077310.109*
C140.6568 (6)0.4888 (4)0.4667 (4)0.092 (2)
H140.6343780.5157850.5096460.111*
C150.7582 (5)0.4568 (3)0.4652 (3)0.0752 (16)
C160.7913 (4)0.4171 (3)0.4014 (3)0.0629 (13)
H160.8599040.3956290.4000210.075*
C210.7416 (4)0.3243 (3)0.1480 (2)0.0518 (11)
C220.6852 (4)0.2672 (3)0.1103 (2)0.0521 (11)
C230.7250 (4)0.2370 (3)0.0428 (3)0.0665 (14)
H230.6870000.1985570.0176200.080*
C240.8208 (5)0.2636 (3)0.0126 (3)0.0719 (15)
H240.8468940.2436980.0330500.086*
C250.8771 (5)0.3193 (3)0.0503 (3)0.0678 (14)
H250.9421060.3366880.0302610.081*
C260.8389 (4)0.3502 (3)0.1175 (3)0.0589 (12)
H260.8779620.3882270.1425230.071*
C310.6682 (5)0.1288 (3)0.2294 (3)0.0813 (17)
H31A0.7195120.1320340.1877600.098*
H31B0.6248220.0831220.2221580.098*
C320.7269 (6)0.1238 (4)0.3034 (4)0.106 (2)
H32A0.7754090.1672630.3081420.127*
H32B0.7698880.0773420.3043470.127*
C330.6496 (6)0.1234 (4)0.3696 (4)0.105 (2)
H33A0.6894780.1229770.4171480.126*
H33B0.6053940.0775390.3676120.126*
C340.5783 (7)0.1938 (4)0.3663 (3)0.108 (2)
H34A0.5251030.1910960.4067970.130*
H34B0.6218800.2392700.3747400.130*
C350.5233 (5)0.2003 (4)0.2927 (3)0.0880 (19)
H35A0.4716690.1587960.2878000.106*
H35B0.4838890.2483380.2910060.106*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.1624 (8)0.1126 (7)0.0671 (4)0.0226 (5)0.0104 (4)0.0245 (4)
S10.0541 (7)0.0761 (9)0.0518 (7)0.0032 (6)0.0116 (5)0.0084 (6)
O10.062 (2)0.096 (3)0.114 (4)0.011 (2)0.016 (2)0.026 (3)
O20.060 (2)0.096 (3)0.077 (2)0.023 (2)0.0044 (18)0.025 (2)
O30.086 (3)0.113 (3)0.066 (2)0.024 (2)0.022 (2)0.010 (2)
N10.060 (2)0.050 (2)0.056 (2)0.0038 (19)0.009 (2)0.0015 (19)
N20.058 (2)0.058 (2)0.054 (2)0.0097 (19)0.0010 (18)0.0104 (18)
C10.058 (3)0.045 (3)0.056 (3)0.009 (2)0.016 (2)0.002 (2)
C110.067 (3)0.038 (3)0.057 (3)0.001 (2)0.021 (2)0.002 (2)
C120.069 (3)0.050 (3)0.088 (4)0.006 (3)0.031 (3)0.012 (3)
C130.076 (4)0.073 (4)0.124 (6)0.010 (3)0.050 (4)0.033 (4)
C140.108 (5)0.070 (4)0.098 (5)0.027 (4)0.056 (4)0.040 (4)
C150.102 (4)0.054 (3)0.069 (3)0.021 (3)0.031 (3)0.013 (3)
C160.087 (4)0.041 (3)0.061 (3)0.001 (3)0.022 (3)0.005 (2)
C210.057 (3)0.054 (3)0.044 (2)0.014 (2)0.007 (2)0.006 (2)
C220.059 (3)0.056 (3)0.041 (2)0.007 (2)0.006 (2)0.008 (2)
C230.084 (4)0.071 (4)0.044 (3)0.006 (3)0.002 (2)0.002 (2)
C240.089 (4)0.080 (4)0.046 (3)0.011 (3)0.015 (3)0.002 (3)
C250.072 (3)0.074 (4)0.057 (3)0.007 (3)0.019 (3)0.010 (3)
C260.065 (3)0.057 (3)0.054 (3)0.002 (2)0.008 (2)0.004 (2)
C310.105 (4)0.067 (4)0.072 (4)0.031 (3)0.010 (3)0.009 (3)
C320.112 (5)0.090 (5)0.115 (5)0.038 (4)0.018 (4)0.033 (4)
C330.147 (6)0.100 (5)0.067 (4)0.023 (5)0.005 (4)0.029 (4)
C340.159 (7)0.104 (5)0.062 (4)0.049 (5)0.010 (4)0.019 (4)
C350.095 (4)0.100 (5)0.069 (4)0.033 (4)0.011 (3)0.018 (3)
Geometric parameters (Å, º) top
Br1—C151.888 (6)C21—C261.395 (7)
S1—O21.428 (4)C22—C231.386 (6)
S1—O31.431 (4)C23—C241.381 (7)
S1—N21.616 (4)C23—H230.9300
S1—C221.767 (5)C24—C251.366 (8)
O1—C121.344 (7)C24—H240.9300
O1—H1A0.8200C25—C261.380 (7)
N1—C11.270 (6)C25—H250.9300
N1—C211.413 (6)C26—H260.9300
N2—C351.451 (6)C31—C321.489 (8)
N2—C311.477 (6)C31—H31A0.9700
C1—C111.449 (6)C31—H31B0.9700
C1—H10.9300C32—C331.506 (9)
C11—C161.382 (7)C32—H32A0.9700
C11—C121.413 (7)C32—H32B0.9700
C12—C131.372 (8)C33—C341.513 (9)
C13—C141.368 (10)C33—H33A0.9700
C13—H130.9300C33—H33B0.9700
C14—C151.377 (9)C34—C351.464 (8)
C14—H140.9300C34—H34A0.9700
C15—C161.376 (7)C34—H34B0.9700
C16—H160.9300C35—H35A0.9700
C21—C221.384 (7)C35—H35B0.9700
O2—S1—O3117.9 (2)C22—C23—H23119.8
O2—S1—N2107.0 (2)C25—C24—C23119.6 (5)
O3—S1—N2111.3 (2)C25—C24—H24120.2
O2—S1—C22109.7 (2)C23—C24—H24120.2
O3—S1—C22107.2 (2)C24—C25—C26121.0 (5)
N2—S1—C22102.8 (2)C24—C25—H25119.5
C12—O1—H1A109.5C26—C25—H25119.5
C1—N1—C21119.7 (4)C25—C26—C21119.8 (5)
C35—N2—C31113.8 (4)C25—C26—H26120.1
C35—N2—S1120.3 (3)C21—C26—H26120.1
C31—N2—S1116.0 (3)N2—C31—C32109.6 (5)
N1—C1—C11121.5 (4)N2—C31—H31A109.7
N1—C1—H1119.2C32—C31—H31A109.7
C11—C1—H1119.2N2—C31—H31B109.7
C16—C11—C12119.6 (4)C32—C31—H31B109.7
C16—C11—C1119.6 (4)H31A—C31—H31B108.2
C12—C11—C1120.7 (5)C31—C32—C33111.0 (6)
O1—C12—C13119.3 (6)C31—C32—H32A109.4
O1—C12—C11122.0 (5)C33—C32—H32A109.4
C13—C12—C11118.7 (6)C31—C32—H32B109.4
C14—C13—C12120.9 (6)C33—C32—H32B109.4
C14—C13—H13119.6H32A—C32—H32B108.0
C12—C13—H13119.6C32—C33—C34109.9 (5)
C13—C14—C15120.9 (5)C32—C33—H33A109.7
C13—C14—H14119.6C34—C33—H33A109.7
C15—C14—H14119.6C32—C33—H33B109.7
C16—C15—C14119.5 (6)C34—C33—H33B109.7
C16—C15—Br1120.9 (5)H33A—C33—H33B108.2
C14—C15—Br1119.6 (4)C35—C34—C33111.6 (6)
C15—C16—C11120.4 (5)C35—C34—H34A109.3
C15—C16—H16119.8C33—C34—H34A109.3
C11—C16—H16119.8C35—C34—H34B109.3
C22—C21—C26119.2 (4)C33—C34—H34B109.3
C22—C21—N1120.8 (4)H34A—C34—H34B108.0
C26—C21—N1120.0 (4)N2—C35—C34111.9 (5)
C21—C22—C23120.0 (4)N2—C35—H35A109.2
C21—C22—S1121.9 (3)C34—C35—H35A109.2
C23—C22—S1118.0 (4)N2—C35—H35B109.2
C24—C23—C22120.4 (5)C34—C35—H35B109.2
C24—C23—H23119.8H35A—C35—H35B107.9
O2—S1—N2—C3530.7 (5)N1—C21—C22—C23176.8 (4)
O3—S1—N2—C3599.4 (5)C26—C21—C22—S1178.2 (3)
C22—S1—N2—C35146.2 (5)N1—C21—C22—S14.4 (6)
O2—S1—N2—C31174.4 (4)O2—S1—C22—C2155.2 (4)
O3—S1—N2—C3144.3 (5)O3—S1—C22—C21175.7 (4)
C22—S1—N2—C3170.2 (4)N2—S1—C22—C2158.3 (4)
C21—N1—C1—C11175.3 (4)O2—S1—C22—C23126.0 (4)
N1—C1—C11—C16178.2 (4)O3—S1—C22—C233.2 (4)
N1—C1—C11—C123.5 (7)N2—S1—C22—C23120.5 (4)
C16—C11—C12—O1180.0 (5)C21—C22—C23—C240.1 (7)
C1—C11—C12—O11.6 (7)S1—C22—C23—C24179.0 (4)
C16—C11—C12—C131.2 (7)C22—C23—C24—C250.9 (8)
C1—C11—C12—C13177.1 (5)C23—C24—C25—C260.9 (8)
O1—C12—C13—C14179.7 (6)C24—C25—C26—C210.2 (8)
C11—C12—C13—C141.4 (9)C22—C21—C26—C250.6 (7)
C12—C13—C14—C150.8 (10)N1—C21—C26—C25176.9 (4)
C13—C14—C15—C160.2 (9)C35—N2—C31—C3255.6 (7)
C13—C14—C15—Br1179.7 (5)S1—N2—C31—C32158.4 (5)
C14—C15—C16—C110.4 (8)N2—C31—C32—C3356.1 (8)
Br1—C15—C16—C11179.9 (4)C31—C32—C33—C3456.3 (9)
C12—C11—C16—C150.2 (7)C32—C33—C34—C3554.6 (9)
C1—C11—C16—C15178.1 (4)C31—N2—C35—C3454.9 (7)
C1—N1—C21—C22135.0 (5)S1—N2—C35—C34160.7 (5)
C1—N1—C21—C2647.6 (6)C33—C34—C35—N253.6 (9)
C26—C21—C22—C230.6 (7)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O1—H1A···N10.821.862.586 (5)146
C1—H1···O2i0.932.543.363 (6)149
C16—H16···O2i0.932.643.461 (6)147
C23—H23···O30.932.432.846 (7)107
C25—H25···O3ii0.932.493.220 (6)135
C26—H26···O1i0.932.623.467 (7)151
C35—H35B···O20.972.432.852 (7)106
Symmetry codes: (i) x+1/2, y, z+1/2; (ii) x+1/2, y+1/2, z.
Global reactivity descriptors of synthesized Schiff bases and standard drugs top
EntryEHOMOELUMOΔEgapIAµXηSω
17-6.2529-1.32874.92426.25291.3287-3.79083.79082.46210.40622.9183
20-5.9851-2.40713.57805.98512.4071-4.19614.19611.78900.55904.9210
21-6.1535-2.00504.14856.15352.0050-4.07934.07932.07430.48214.0112
18-6.1546-2.00504.14966.15462.0050-4.07984.07982.07480.48204.0112
19-6.5745-2.49634.07826.57452.4963-4.53544.53542.03910.49045.0439
22-6.1778-1.60354.57436.17781.6035-3.89073.89072.28720.43723.3092
23-6.6167-2.27134.34546.61672.2713-4.44404.44402.17270.46034.5448
5-Flu-6.7827-1.26595.51686.78271.2659-4.02434.02432.75840.36252.9355
Cap-6.4578-1.60194.85596.45781.6019-4.02994.02992.42790.41193.3444
Tri-6.9868-1.37255.61436.98681.3725-4.17974.17972.80710.35623.1117
Where ΔEgap is energy gap or charge separation, I is ionization potential, A is electron affinity, µ is chemical potential, X is electronegativity, η is global hardness, S is global softness and ω is electrophilicity index. Tri is trifluridine, Cap is capecitabine and 5-Flu is 5-fluorouracil.
Summary of the binding energy of Schiff bases with poly(ADP-ribose) polymerase top
Optimized Schiff basesSummarized drug-likeness and toxicity6kro
6kro_23_E=714.59mildly nondrug-like and toxic-11.1
6kro_22_E=797.81nontoxic but mildly nondrug-like-10.3
6kro_21_E=687.32nondrug-like and nontoxic-9.9
6kro_20_E=635.91drug-like and nontoxic-9.5
6kro_17_E=666.05drug-like and nontoxic-9.2
6kro_18_E=685.47drug-like and nontoxic-8.7
6kro_19_E=748.77drug-like but mildly toxic-6.7
6kro_trifluridine_E=282.80drug-like but mildly toxic-8
6kro_capecitabine_E=624.15drug-like but highly toxic-7.9
6kro_5-fluorouracil_E=45.84mildly nondrug-like and toxic-5.5
In-silico toxicity study and drug-likeness of 20 using ProTox-II and SwissADME top
Druglikeness
Toxicity
TargetPredictionProbability
HepatotoxicityInactive0.57
CarcinogenicityInactive0.57
ImmunotoxicityInactive0.87
MutagenicityInactive0.71
CytotoxicityInactive0.73
In-silico toxicity study and drug-likeness of 23 using ProTox-II and SwissADME top
Druglikeness
Toxicity
TargetPredictionProbability
HepatotoxicityInactive0.62
CarcinogenicityInactive0.57
ImmunotoxicityInactive0.94
MutagenicityActive0.68
CytotoxicityInactive0.78
In-silico toxicity study and drug-likeness of trifluridine top
Druglikeness
Toxicity
TargetPredictionProbability
HepatotoxicityInactive0.76
CarcinogenicityInactive0.60
ImmunotoxicityInactive0.99
MutagenicityActive0.64
CytotoxicityInactive0.88
In-silico toxicity study and drug-likeness of 5-fluorouracil top
Druglikeness
Toxicity
TargetPredictionProbability
HepatotoxicityInactive0.78
CarcinogenicityActive0.85
ImmunotoxicityInactive0.99
MutagenicityInactive0.88
CytotoxicityInactive0.93
Physicochemical properties of synthesized Schiff bases and reference drugs top
CompoundMrNo. of heavy atomsFraction CCp3No. rotational bondsNo. hydrogen-bond acceptorsNo. hydrogen-bond donorsTPSAlog Kp (cm s-1)Bioavailability score
17344.43240.2845178.35-6.370.55
20409.3240.2445178.35-6.530.55
21457.34280.144178.35-6.040.55
18423.32250.2845178.35-6.360.55
19389.43270.28571124.17-6.770.55
22471.37290.1445178.35-6.170.55
23437.47310.14571124.17-6.580.55
Trifluridine296.2200.6383104.55-8.430.55
Capecitabine359.35250.67883122.91-8.090.55
5-Fluorouracil130.089003265.72-7.730.55
Druglikeness rule violations, in-silico toxicity study of trifluridine and 5-fluorouracil for comparison top
Druglikeness rules' violationsBlood–brain distribution and metabolism
CompoundsLipinskiGhoseVeberEganMueggeGI absorptionBBB permeant
1700000HighNo
2000000HighNo
2100000HighNo
1800000HighNo
1900000HighNo
2200000LowNo
2300000HighNo
Trifluridine00000HighNo
Capecitabine00000HighNo
5-Fluorouracil00000HighNo
 

Acknowledgements

This project was graciously funded by the University of Lagos Central Research Committee, the Nigerian Government's TetFund IBR and the National Research Foundation (NRF) of South Africa. We extend our gratitude to the Center for High Performance Computing (CHPC) in Cape Town, South Africa, for furnishing the necessary computational resources on the Schrödinger Platform that facilitated our mol­ecular modelling studies focused on protein preparation. It is important to note that the authors declare no financial or non-financial conflicts of inter­est related to this study. The conceptual framework and design of the research were collaboratively developed by all contributing authors. Material preparation, data collection and analysis were per­formed by Sherif O. Kolade, Eric C. Hosten, Allen T. Gordon, Idris A. Olasupo and Olayinka T. Asekun. The first draft of the manuscript was written by Sherif O. Kolade, Adeniyi S. Ogunlaja and Oluwole B. Familoni, and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

Funding information

The following funding is acknowledged: University of Lagos Central Research Committee (grant No. CRC 2015/25 to Oluwole Familoni); Nigerian Government TETFund IBR (grant No. CRC/TETFUND 2018/016 to Oluwole Familoni); National Research Foundation (NRF) of South Africa (grant No. 129887 to Adeniyi Ogunlaja).

References

First citationAbdelsalam, M. M., Bedair, M. A., Hassan, A. M., Heakal, B. H., Younis, A., Elbialy, Z. I., Badawy, M. A., Fawzy, H. E. & Fareed, S. A. (2022). Arabian J. Chem. 15, 103491.  CrossRef Google Scholar
First citationAbd El-Wahab, H., Abd El-Fattah, M., El-alfy, H. M. Z., Owda, M. E., Lin, L. & Hamdy, I. (2020). Prog. Org. Coat. 142, 105577.  Google Scholar
First citationAbd-Elzaher, M. M., Labib, A. A., Mousa, H. A., Moustafa, S. A., Ali, M. M. & El-Rashedy, A. A. (2016). Beni-Suef Univ. J. Basic Appl. Sci. 5, 85–96.  Google Scholar
First citationAfsan, Z., Roisnel, T., Tabassum, S. & Arjmand, F. (2020). Bioorg. Chem. 94, 103427.  CrossRef PubMed Google Scholar
First citationAkman, F. (2019). Cellul. Chem. Technol. 53, 243–250.  CrossRef CAS Google Scholar
First citationAlblewi, F. F., Alsehli, M. H., Hritani, Z. M., Eskandrani, A., Alsaedi, W. H., Alawad, M. O., Elhenawy, A. A., Ahmed, H. Y., El-Gaby, M. S. A., Afifi, T. H. & Okasha, R. M. (2023). Int. J. Mol. Sci. 24, 16716.  CrossRef PubMed Google Scholar
First citationAltürk, S., Tamer, Ö., Avcı, D. & Atalay, Y. (2015). J. Organomet. Chem. 797, 110–119.  Google Scholar
First citationBecke, A. D. (1993). J. Chem. Phys. 98, 5648–5652.  CrossRef CAS Web of Science Google Scholar
First citationBehrens, J. (2000). Ann. N. Y. Acad. Sci. 910, 21–35.  CrossRef PubMed CAS Google Scholar
First citationBruker (2016). APEX2, SAINT, SADABS and TWINABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationChoudhary, N., Bee, S., Gupta, A. & Tandon, P. (2013). Comput. Theor. Chem. 1016, 8–21.  CrossRef CAS Google Scholar
First citationDassault Systémes (2020). Discovery Studio. BIOVIA, San Diego, CA, USA. https://www.3ds.com/Google Scholar
First citationDueke-Eze, C. U., Fasina, T. M., Oluwalana, A. E., Familoni, O. B., Mphalele, J. M. & Onubuogu, C. (2020). Sci. Afr. 9, e00522.  Google Scholar
First citationEisemann, T. & Pascal, J. M. (2020). Cell. Mol. Life Sci. 77, 19–33.  CrossRef CAS PubMed Google Scholar
First citationElsamra, R. M., Masoud, M. S. & Ramadan, A. M. (2022). Sci. Rep. 12, 20192.  CrossRef PubMed Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFeng, X. & Koh, D. W. (2013). Int. Rev. Cell Mol. Biol. 304, 227–281.  CrossRef CAS PubMed Google Scholar
First citationFriesner, R. A., Murphy, R. B., Repasky, M. P., Frye, L. L., Greenwood, J. R., Halgren, T. A., Sanschagrin, P. C. & Mainz, D. T. (2006). J. Med. Chem. 49, 6177–6196.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGao, C., Xiao, G. & Hu, J. (2014). Cell Biosci. 4, 13.  CrossRef PubMed Google Scholar
First citationGordon, A. T., Hosten, E. C. & Ogunlaja, A. S. (2022). Pharmaceuticals, 15, 1240.  CrossRef PubMed Google Scholar
First citationHall, M. L., Goldfeld, D. A., Bochevarov, A. D. & Friesner, R. A. (2009). J. Chem. Theory Comput. 5, 2996–3009.  CrossRef CAS PubMed Google Scholar
First citationHuang, H. & He, X. (2008). Curr. Opin. Cell Biol. 20, 119–125.  CrossRef PubMed CAS Google Scholar
First citationHübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281–1284.  Web of Science CrossRef IUCr Journals Google Scholar
First citationIrfan, A., Rubab, L., Rehman, M. U., Anjum, R., Ullah, S., Marjana, M., Qadeer, S. & Sana, S. (2020). Heterocycl. Commun. 26, 46–59.  CrossRef CAS Google Scholar
First citationJędrzejczyk, M., Janczak, J. & Huczyński, A. (2022). J. Mol. Struct. 1263, 133129.  Google Scholar
First citationKolade, S. O., Izunobi, J. U., Gordon, A. T., Hosten, E. C., Olasupo, I. A., Ogunlaja, A. S., Asekun, O. T. & Familoni, O. B. (2022). Acta Cryst. C78, 730–742.  CrossRef IUCr Journals Google Scholar
First citationKolade, S. O., Izunobi, J. U., Hosten, E. C., Olasupo, I. A., Ogunlaja, A. S. & Familoni, O. B. (2020). Acta Cryst. C76, 810–820.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationKondapuram, S. K., Sarvagalla, S. & Coumar, M. S. (2021). In Molecular Docking for Comput-Aided Drug Design, pp. 463–477. New York: Academic Press.  Google Scholar
First citationKumar, A., Bhagat, K. K., Singh, A. K., Singh, H., Angre, T., Verma, A., Khalilullah, H., Jaremko, M., Emwas, A. H. & Kumar, P. (2023a). RSC Adv. 13, 6872–6908.  CrossRef CAS PubMed Google Scholar
First citationKumar, A., Singh, A. K., Singh, H., Vijayan, V., Kumar, D., Naik, J. & Kumar, P. (2023b). Pharmaceuticals, 16, 299.  CrossRef PubMed Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMahmood, A., Irfan, A. & Wang, J. L. (2022). Chem. Eur. J. 28, e202103712.  CrossRef PubMed Google Scholar
First citationMatela, G. (2020). Anticancer Agents Med. Chem. 20, 1908–1917.  CrossRef CAS PubMed Google Scholar
First citationMeyer, R. G., Meyer-Ficca, M. L., Jacobson, E. L. & Jacobson, M. K. (2006). Enzymes in poly(ADP-ribose) metabolism, edited by A. Bürkle, pp. 1–12. Georgetown: Landes Bioscience.  Google Scholar
First citationMuhammad-Ali, M. A., Jasim, E. Q. & Al-Saadoon, A. H. (2023). J. Med. Chem. Sci. 6, 2128–2139.  CAS Google Scholar
First citationPai, S. G., Carneiro, B. A., Mota, J. M., Costa, R., Leite, C. A., Barroso-Sousa, R. & Giles, F. J. (2017). J. Hematol. Oncol. 10, 101.  CrossRef PubMed Google Scholar
First citationPandi, S., Kulanthaivel, L., Subbaraj, G. K., Rajaram, S. & Subramanian, S. (2022). BioMed Res. Int. 2022, 3338549.  CrossRef PubMed Google Scholar
First citationPawar, S. S. & Rohane, S. H. (2021). Asian J. Res. Chem. 14, 86-–88.  CrossRef Google Scholar
First citationPereira, F., Xiao, K., Latino, D. A., Wu, C., Zhang, Q. & Aires-de-Sousa, J. (2017). J. Chem. Inf. Model. 57, 11–21.  CrossRef CAS PubMed Google Scholar
First citationProkopenko, Y. S., Perekhoda, L. O. & Georgiyants, V. A. (2019). J. Appl. Pharm. Sci. 9, 066–072.  CAS Google Scholar
First citationSalehi, M., Kubicki, M., Galini, M., Jafari, M. & Malekshah, R. E. (2019). J. Mol. Struct. 1180, 595–602.  CrossRef CAS 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 citationShirai, F., Mizutani, A., Yashiroda, Y., Tsumura, T., Kano, Y., Muramatsu, Y., Chikada, T., Yuki, H., Niwa, H., Sato, S., Washizuka, K., Koda, Y., Mazaki, Y., Jang, M. K., Yoshida, H., Nagamori, A., Okue, M., Watanabe, T., Kitamura, K., Shitara, E., Honma, T., Umehara, T., Shirouzu, M., Fukami, T., Seimiya, H., Yoshida, M. & Koyama, H. (2020). J. Med. Chem. 63, 4183–4204.  CrossRef CAS PubMed Google Scholar
First citationSilva, R. R. da, Ramalho, T. C., Santos, J. M. & Figueroa-Villar, J. D. (2006). J. Phys. Chem. A, 110, 1031–1040.  PubMed 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 citationSpek, A. L. (2020). Acta Cryst. E76, 1–11.  Web of Science CrossRef IUCr Journals Google Scholar
First citationTrott, O. & Olson, A. J. (2010). J. Comput. Chem. 31, 455–461.  Web of Science CrossRef PubMed CAS Google Scholar
First citationYele, V., Sigalapalli, D. K., Jupudi, S. & Mohammed, A. A. (2021). J. Mol. Model. 27, 359.  CrossRef 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 logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296
Follow Acta Cryst. C
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds