research papers
Synthesis, in-silico evaluation of arylsulfonamide for potential activity against colon cancer
andaDepartment 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
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 to elucidate their pharmacokinetic profiles and potential. Impressively, the ADMET analysis showcased encouraging drug-likeness properties of the newly synthesized 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 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.
Keywords: Schiff base; arylsulfonamide; colon cancer; crystal structure; molecular docking; ADMET.
CCDC reference: 2305610
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,b). 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 have emerged as a class of compounds with significant potential due to their versatile chemical structures and promising pharmacological profiles (Irfan et al., 2020; Muhammad-Ali et al., 2023; Dueke-Eze et al., 2020).
Schiff bases, characterized by their imine et al., 2023). 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; Elsamra et al., 2022). The rationale behind this hypothesis centres on exploring the anticancer potential of particularly focusing on their interaction with colon cancer. Extensive literature has already underscored their efficacy as anticancer agents (Abd-Elzaher et al., 2016). Remarkably, have demonstrated activity against colon cancer and have been documented for such effects (Matela, 2020). Notably, the combination of benzenesulfonamide with a Schiff base has been reported, merging the bioactive attributes of sulfonamides and to investigate potential synergies between these well-established functional groups (Afsan et al., 2020). The cumulative evidence of their enhanced activity spurred our interest in undertaking the present study.
(–C=N–), have been studied extensively for their diverse pharmacological activities, including anticancer properties (AlblewiTankyrase 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; Eisemann & Pascal, 2020). 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). 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). Under normal circumstances, Axin aids in regulating the Wnt pathway by facilitating β-catenin degradation (Huang & He, 2008). 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). Inhibiting specific amino acid residues in human tankyrase poly(ADP-ribose) polymerase using such as (E)-2-[(2-hydroxybenzylidene)amino]benzenesulfonamide derivatives (17–23) (Fig. 1), could potentially prevent the accumulation of β-catenin, holding promise for effective intervention (Meyer et al., 2006).
Our research also delves into the et al., 2020), which, along with our exploration of the of a Schiff base sulfonamide, forms a comprehensive investigation into the potential therapeutic avenues these compounds may offer.
of benzenesulfonamides (Kolade2. 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 n-hexane (2 or 1.4:1 v/v) solvent system visualized under a UV lamp (254 nm). 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. 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.
(TLC) on Merck silica gel 60 F254 precoated plates using a dichloromethane/2.2. Synthesis and crystallization
2.2.1. Synthesis of sulfonamide Schiff bases
The general reaction scheme for the formation of potentially bioactive sulfonamide . The N-cycloamino-2-sulfanilamides were prepared as reported previously (Kolade et al., 2022) by reacting the aminosulfanilamides with substituted o-salicylaldehyde either at room temperature or under reflux to obtain the required 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.
is shown in Fig. 12.2.1.1. 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 18–23. To a stirring solution of N-cycloamino-2-sulfanilamides 10–13 (1.0 mmol) in ethanol (10 ml) was added 5-bromo(nitro)-o-salicylaldehydes 15–16 (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 18–23 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), its protein (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). Energy minimization was carried out, followed by conversion into protein databank partial charge (pdbqt) ligands. The 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). 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).
2.3.3. Molecular docking
Docking simulations were performed with PyRx AutoDock using the Lamarkian 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). 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 . 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).
details are summarized in Table 13. Results and discussion
3.1. Chemistry
N1-(51-Substituted-21-hydroxybenzylidenyl)-N-cycloamino-2-sulfanilamides 17–23 were prepared from the reaction of N-cycloamino-2-sulfanilamides 10–13 [as reported previously by Kolade et al. (2022)] with substituted o-salicylaldehyde either at room temperature or under reflux to obtain the required 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 17–23 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). 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 (ligands) 17–23 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 is corroborated by the 13C NMR spectra (see supporting information) of compounds 17–23, 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 Pbca (Table 1). 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). 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.
|
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; Prokopenko et al., 2019; Jędrzejczyk et al., 2022; Pandi et al., 2022). 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), (A), chemical potential (μ), (χ), global hardness (η), global softness (S) and global (ω) values (Gordon et al., 2022).
3.3.2. Global reactivity descriptors of the synthesized 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). 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).
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). 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). Considering the context of the LUMO within the synthesized 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).
|
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). 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), π–π 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 π–π 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).
Similarly, Schiff base 21 (Fig. 3) 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) 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 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).
Within the structures of 5-fluorouracil and trifluridine (Fig. 6), the HOMO and LUMO exhibit localization in distinct regions of each respective molecule. This localization directly signifies the occurrence of charge-transfer processes.
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). 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; Akman, 2019; Choudhary et al., 2013; Abdelsalam et al., 2022). 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 The (μ) spans from −4.5354 eV (lowest) for 19 to −3.7908 eV (highest) for 17. 19 and 20 exhibit the highest index (ω), with values of 5.044 and 4.9209 eV, respectively. In contrast, Schiff base 17 showcases the lowest 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.
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; Friesner et al., 2006).
Upon scrutinizing the MESP mappings of individual compounds, distinct patterns emerge. Schiff base 17 [Fig. 7(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(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(a)], an evident gradient from green to blue characterizes the MESP map. Analogously, the MESP portrayal of 18 [Fig. 9(b)] reflects this trend, except for the S=O functional-group regions which assume a red hue. Both 19 [Fig. 9(a)] and 22 [Fig. 9(b)] display a blending of blue and green regions in their respective MESP renderings. For 23 [Fig. 10(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(b)], predominantly depicted in blue in its MESP representation. The standard drug 5-fluorouracil [Fig. 11(a)] showcases the entire spectrum of colour variations across its surface. Similarly, trifluridine [Fig. 11(b)] transitions from blue to red, thereby illustrating its surface characteristics encompassing the 2-(hydroxymethyl)tetrahydrofuran-3-ol moiety.
3.4. Docking studies
3.4.1. 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 (the structures are presented in Fig. 12). In addition, the drug-likeness and toxicity of the synthesized and reference drugs were also investigated (Tables 5–10).
|
|
|
|
|
|
|
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 (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).
The 2D and 3D structures of 20 showing the interacting amino acid residues [Fig. 13(a)], bond lengths [Fig. 13(b)], hydrophobic interactions [Fig. 13(c)] and solvent-accessibility surface [Fig. 13(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(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.
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(c)]. While Arg1100 had good solvent-accessibility surface interactions, other interacting residues had excellent solvent interactions with ligand 20 [Figs. 13(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(a) and 14(b).
Noticeably, the solvent-accessibility surface of the reference drug seems better, as all interacting amino acid residues interact in the blue region [Fig. 14(d)]; however, it has comparably lower binding energy (Table 4), i.e. poorer hydrophobicity than exhibited by most drug-like candidates [Fig. 14(c)], and it possesses mutagenic toxicity (Table 7).
3.4.2. Toxicity and drug-likeness of 17–23 and reference drugs
The SMILES (simplified molecular-input line-entry system) of the synthesized 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. 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). 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).
and the reference drugs were obtainedIn 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 and 9). Unlike synthesized 17, 20 and 18, this study also reveals that 5-fluorouracil fails some drug-likeness tests in addition to its toxic nature (Table 9).
In the course of the drug-likeness investigation, physicochemical parameters and drug-likeness violations of the and 10. 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 are blood–brain barrier permeant, making them safe without any unwarranted interference with the central nervous system (Table 9).
and the reference drugs trifluridine and 5-fluorouracil were also compared, as shown in Tables 94. Conclusion
The successful synthesis, characterization and analysis of the intermolecular interactions of N1-(51-substituted-21-hydroxybenzylidenyl)-N-cycloamino-2-sulfanilamides (compounds 17–23) 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 π–π stacking play a crucial role in the molecular cohesion of these compounds. Furthermore, the DFT results, when combined with molecular reveal that the and 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 17–23.
Supporting information
CCDC reference: 2305610
https://doi.org/10.1107/S205322962400233X/ef3052sup1.cif
contains datablocks I, global. DOI:Structure factors: contains datablock I. DOI: https://doi.org/10.1107/S205322962400233X/ef3052Isup2.hkl
Supporting information file. DOI: https://doi.org/10.1107/S205322962400233X/ef3052sup3.pdf
C18H19BrN2O3S | Dx = 1.484 Mg m−3 |
Mr = 423.32 | Melting point: 427.45 K |
Orthorhombic, Pbca | Mo Kα radiation, λ = 0.71073 Å |
a = 12.4070 (9) Å | Cell parameters from 9105 reflections |
b = 17.4250 (14) Å | θ = 2.3–24.5° |
c = 17.5276 (12) Å | µ = 2.30 mm−1 |
V = 3789.3 (5) Å3 | T = 296 K |
Z = 8 | Platelet, yellow |
F(000) = 1728 | 0.84 × 0.43 × 0.12 mm |
Bruker APEXII CCD diffractometer | 3356 independent reflections |
Radiation source: sealed tube | 2255 reflections with I > 2σ(I) |
Graphite monochromator | Rint = 0.060 |
Detector resolution: 8.3333 pixels mm-1 | θmax = 25.1°, θmin = 2.3° |
φ and ω scans | h = −14→14 |
Absorption correction: multi-scan (SADABS; Bruker, 2016) | k = −19→20 |
Tmin = 0.133, Tmax = 0.241 | l = −20→20 |
27860 measured reflections |
Refinement on F2 | Primary atom site location: dual |
Least-squares matrix: full | Secondary atom site location: difference Fourier map |
R[F2 > 2σ(F2)] = 0.062 | Hydrogen site location: inferred from neighbouring sites |
wR(F2) = 0.172 | H-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 parameters | Δρmax = 0.52 e Å−3 |
0 restraints | Δρmin = −0.37 e Å−3 |
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). |
x | y | z | Uiso*/Ueq | ||
Br1 | 0.85040 (8) | 0.46864 (5) | 0.55005 (4) | 0.1140 (4) | |
S1 | 0.56327 (10) | 0.22940 (8) | 0.14642 (7) | 0.0607 (4) | |
O1 | 0.5496 (3) | 0.4349 (3) | 0.2834 (3) | 0.0907 (13) | |
H1A | 0.579549 | 0.413541 | 0.247549 | 0.136* | |
O2 | 0.4882 (3) | 0.2902 (2) | 0.1591 (2) | 0.0777 (11) | |
O3 | 0.5316 (3) | 0.1682 (2) | 0.0968 (2) | 0.0883 (12) | |
N1 | 0.6996 (3) | 0.3585 (2) | 0.2148 (2) | 0.0555 (10) | |
N2 | 0.5983 (3) | 0.1974 (2) | 0.2293 (2) | 0.0565 (10) | |
C1 | 0.7606 (4) | 0.3689 (3) | 0.2721 (3) | 0.0533 (11) | |
H1 | 0.830912 | 0.350473 | 0.270688 | 0.064* | |
C11 | 0.7232 (4) | 0.4088 (3) | 0.3396 (3) | 0.0543 (12) | |
C12 | 0.6188 (5) | 0.4412 (3) | 0.3421 (3) | 0.0688 (15) | |
C13 | 0.5885 (5) | 0.4815 (3) | 0.4058 (4) | 0.091 (2) | |
H13 | 0.520618 | 0.504049 | 0.407731 | 0.109* | |
C14 | 0.6568 (6) | 0.4888 (4) | 0.4667 (4) | 0.092 (2) | |
H14 | 0.634378 | 0.515785 | 0.509646 | 0.111* | |
C15 | 0.7582 (5) | 0.4568 (3) | 0.4652 (3) | 0.0752 (16) | |
C16 | 0.7913 (4) | 0.4171 (3) | 0.4014 (3) | 0.0629 (13) | |
H16 | 0.859904 | 0.395629 | 0.400021 | 0.075* | |
C21 | 0.7416 (4) | 0.3243 (3) | 0.1480 (2) | 0.0518 (11) | |
C22 | 0.6852 (4) | 0.2672 (3) | 0.1103 (2) | 0.0521 (11) | |
C23 | 0.7250 (4) | 0.2370 (3) | 0.0428 (3) | 0.0665 (14) | |
H23 | 0.687000 | 0.198557 | 0.017620 | 0.080* | |
C24 | 0.8208 (5) | 0.2636 (3) | 0.0126 (3) | 0.0719 (15) | |
H24 | 0.846894 | 0.243698 | −0.033050 | 0.086* | |
C25 | 0.8771 (5) | 0.3193 (3) | 0.0503 (3) | 0.0678 (14) | |
H25 | 0.942106 | 0.336688 | 0.030261 | 0.081* | |
C26 | 0.8389 (4) | 0.3502 (3) | 0.1175 (3) | 0.0589 (12) | |
H26 | 0.877962 | 0.388227 | 0.142523 | 0.071* | |
C31 | 0.6682 (5) | 0.1288 (3) | 0.2294 (3) | 0.0813 (17) | |
H31A | 0.719512 | 0.132034 | 0.187760 | 0.098* | |
H31B | 0.624822 | 0.083122 | 0.222158 | 0.098* | |
C32 | 0.7269 (6) | 0.1238 (4) | 0.3034 (4) | 0.106 (2) | |
H32A | 0.775409 | 0.167263 | 0.308142 | 0.127* | |
H32B | 0.769888 | 0.077342 | 0.304347 | 0.127* | |
C33 | 0.6496 (6) | 0.1234 (4) | 0.3696 (4) | 0.105 (2) | |
H33A | 0.689478 | 0.122977 | 0.417148 | 0.126* | |
H33B | 0.605394 | 0.077539 | 0.367612 | 0.126* | |
C34 | 0.5783 (7) | 0.1938 (4) | 0.3663 (3) | 0.108 (2) | |
H34A | 0.525103 | 0.191096 | 0.406797 | 0.130* | |
H34B | 0.621880 | 0.239270 | 0.374740 | 0.130* | |
C35 | 0.5233 (5) | 0.2003 (4) | 0.2927 (3) | 0.0880 (19) | |
H35A | 0.471669 | 0.158796 | 0.287800 | 0.106* | |
H35B | 0.483889 | 0.248338 | 0.291006 | 0.106* |
U11 | U22 | U33 | U12 | U13 | U23 | |
Br1 | 0.1624 (8) | 0.1126 (7) | 0.0671 (4) | −0.0226 (5) | 0.0104 (4) | −0.0245 (4) |
S1 | 0.0541 (7) | 0.0761 (9) | 0.0518 (7) | 0.0032 (6) | −0.0116 (5) | 0.0084 (6) |
O1 | 0.062 (2) | 0.096 (3) | 0.114 (4) | 0.011 (2) | 0.016 (2) | −0.026 (3) |
O2 | 0.060 (2) | 0.096 (3) | 0.077 (2) | 0.023 (2) | −0.0044 (18) | 0.025 (2) |
O3 | 0.086 (3) | 0.113 (3) | 0.066 (2) | −0.024 (2) | −0.022 (2) | −0.010 (2) |
N1 | 0.060 (2) | 0.050 (2) | 0.056 (2) | 0.0038 (19) | 0.009 (2) | −0.0015 (19) |
N2 | 0.058 (2) | 0.058 (2) | 0.054 (2) | 0.0097 (19) | 0.0010 (18) | 0.0104 (18) |
C1 | 0.058 (3) | 0.045 (3) | 0.056 (3) | 0.009 (2) | 0.016 (2) | −0.002 (2) |
C11 | 0.067 (3) | 0.038 (3) | 0.057 (3) | −0.001 (2) | 0.021 (2) | −0.002 (2) |
C12 | 0.069 (3) | 0.050 (3) | 0.088 (4) | −0.006 (3) | 0.031 (3) | −0.012 (3) |
C13 | 0.076 (4) | 0.073 (4) | 0.124 (6) | −0.010 (3) | 0.050 (4) | −0.033 (4) |
C14 | 0.108 (5) | 0.070 (4) | 0.098 (5) | −0.027 (4) | 0.056 (4) | −0.040 (4) |
C15 | 0.102 (4) | 0.054 (3) | 0.069 (3) | −0.021 (3) | 0.031 (3) | −0.013 (3) |
C16 | 0.087 (4) | 0.041 (3) | 0.061 (3) | −0.001 (3) | 0.022 (3) | −0.005 (2) |
C21 | 0.057 (3) | 0.054 (3) | 0.044 (2) | 0.014 (2) | 0.007 (2) | 0.006 (2) |
C22 | 0.059 (3) | 0.056 (3) | 0.041 (2) | 0.007 (2) | −0.006 (2) | 0.008 (2) |
C23 | 0.084 (4) | 0.071 (4) | 0.044 (3) | 0.006 (3) | −0.002 (2) | −0.002 (2) |
C24 | 0.089 (4) | 0.080 (4) | 0.046 (3) | 0.011 (3) | 0.015 (3) | 0.002 (3) |
C25 | 0.072 (3) | 0.074 (4) | 0.057 (3) | 0.007 (3) | 0.019 (3) | 0.010 (3) |
C26 | 0.065 (3) | 0.057 (3) | 0.054 (3) | 0.002 (2) | 0.008 (2) | 0.004 (2) |
C31 | 0.105 (4) | 0.067 (4) | 0.072 (4) | 0.031 (3) | 0.010 (3) | 0.009 (3) |
C32 | 0.112 (5) | 0.090 (5) | 0.115 (5) | 0.038 (4) | −0.018 (4) | 0.033 (4) |
C33 | 0.147 (6) | 0.100 (5) | 0.067 (4) | 0.023 (5) | −0.005 (4) | 0.029 (4) |
C34 | 0.159 (7) | 0.104 (5) | 0.062 (4) | 0.049 (5) | 0.010 (4) | 0.019 (4) |
C35 | 0.095 (4) | 0.100 (5) | 0.069 (4) | 0.033 (4) | 0.011 (3) | 0.018 (3) |
Br1—C15 | 1.888 (6) | C21—C26 | 1.395 (7) |
S1—O2 | 1.428 (4) | C22—C23 | 1.386 (6) |
S1—O3 | 1.431 (4) | C23—C24 | 1.381 (7) |
S1—N2 | 1.616 (4) | C23—H23 | 0.9300 |
S1—C22 | 1.767 (5) | C24—C25 | 1.366 (8) |
O1—C12 | 1.344 (7) | C24—H24 | 0.9300 |
O1—H1A | 0.8200 | C25—C26 | 1.380 (7) |
N1—C1 | 1.270 (6) | C25—H25 | 0.9300 |
N1—C21 | 1.413 (6) | C26—H26 | 0.9300 |
N2—C35 | 1.451 (6) | C31—C32 | 1.489 (8) |
N2—C31 | 1.477 (6) | C31—H31A | 0.9700 |
C1—C11 | 1.449 (6) | C31—H31B | 0.9700 |
C1—H1 | 0.9300 | C32—C33 | 1.506 (9) |
C11—C16 | 1.382 (7) | C32—H32A | 0.9700 |
C11—C12 | 1.413 (7) | C32—H32B | 0.9700 |
C12—C13 | 1.372 (8) | C33—C34 | 1.513 (9) |
C13—C14 | 1.368 (10) | C33—H33A | 0.9700 |
C13—H13 | 0.9300 | C33—H33B | 0.9700 |
C14—C15 | 1.377 (9) | C34—C35 | 1.464 (8) |
C14—H14 | 0.9300 | C34—H34A | 0.9700 |
C15—C16 | 1.376 (7) | C34—H34B | 0.9700 |
C16—H16 | 0.9300 | C35—H35A | 0.9700 |
C21—C22 | 1.384 (7) | C35—H35B | 0.9700 |
O2—S1—O3 | 117.9 (2) | C22—C23—H23 | 119.8 |
O2—S1—N2 | 107.0 (2) | C25—C24—C23 | 119.6 (5) |
O3—S1—N2 | 111.3 (2) | C25—C24—H24 | 120.2 |
O2—S1—C22 | 109.7 (2) | C23—C24—H24 | 120.2 |
O3—S1—C22 | 107.2 (2) | C24—C25—C26 | 121.0 (5) |
N2—S1—C22 | 102.8 (2) | C24—C25—H25 | 119.5 |
C12—O1—H1A | 109.5 | C26—C25—H25 | 119.5 |
C1—N1—C21 | 119.7 (4) | C25—C26—C21 | 119.8 (5) |
C35—N2—C31 | 113.8 (4) | C25—C26—H26 | 120.1 |
C35—N2—S1 | 120.3 (3) | C21—C26—H26 | 120.1 |
C31—N2—S1 | 116.0 (3) | N2—C31—C32 | 109.6 (5) |
N1—C1—C11 | 121.5 (4) | N2—C31—H31A | 109.7 |
N1—C1—H1 | 119.2 | C32—C31—H31A | 109.7 |
C11—C1—H1 | 119.2 | N2—C31—H31B | 109.7 |
C16—C11—C12 | 119.6 (4) | C32—C31—H31B | 109.7 |
C16—C11—C1 | 119.6 (4) | H31A—C31—H31B | 108.2 |
C12—C11—C1 | 120.7 (5) | C31—C32—C33 | 111.0 (6) |
O1—C12—C13 | 119.3 (6) | C31—C32—H32A | 109.4 |
O1—C12—C11 | 122.0 (5) | C33—C32—H32A | 109.4 |
C13—C12—C11 | 118.7 (6) | C31—C32—H32B | 109.4 |
C14—C13—C12 | 120.9 (6) | C33—C32—H32B | 109.4 |
C14—C13—H13 | 119.6 | H32A—C32—H32B | 108.0 |
C12—C13—H13 | 119.6 | C32—C33—C34 | 109.9 (5) |
C13—C14—C15 | 120.9 (5) | C32—C33—H33A | 109.7 |
C13—C14—H14 | 119.6 | C34—C33—H33A | 109.7 |
C15—C14—H14 | 119.6 | C32—C33—H33B | 109.7 |
C16—C15—C14 | 119.5 (6) | C34—C33—H33B | 109.7 |
C16—C15—Br1 | 120.9 (5) | H33A—C33—H33B | 108.2 |
C14—C15—Br1 | 119.6 (4) | C35—C34—C33 | 111.6 (6) |
C15—C16—C11 | 120.4 (5) | C35—C34—H34A | 109.3 |
C15—C16—H16 | 119.8 | C33—C34—H34A | 109.3 |
C11—C16—H16 | 119.8 | C35—C34—H34B | 109.3 |
C22—C21—C26 | 119.2 (4) | C33—C34—H34B | 109.3 |
C22—C21—N1 | 120.8 (4) | H34A—C34—H34B | 108.0 |
C26—C21—N1 | 120.0 (4) | N2—C35—C34 | 111.9 (5) |
C21—C22—C23 | 120.0 (4) | N2—C35—H35A | 109.2 |
C21—C22—S1 | 121.9 (3) | C34—C35—H35A | 109.2 |
C23—C22—S1 | 118.0 (4) | N2—C35—H35B | 109.2 |
C24—C23—C22 | 120.4 (5) | C34—C35—H35B | 109.2 |
C24—C23—H23 | 119.8 | H35A—C35—H35B | 107.9 |
O2—S1—N2—C35 | 30.7 (5) | N1—C21—C22—C23 | 176.8 (4) |
O3—S1—N2—C35 | −99.4 (5) | C26—C21—C22—S1 | 178.2 (3) |
C22—S1—N2—C35 | 146.2 (5) | N1—C21—C22—S1 | −4.4 (6) |
O2—S1—N2—C31 | 174.4 (4) | O2—S1—C22—C21 | 55.2 (4) |
O3—S1—N2—C31 | 44.3 (5) | O3—S1—C22—C21 | −175.7 (4) |
C22—S1—N2—C31 | −70.2 (4) | N2—S1—C22—C21 | −58.3 (4) |
C21—N1—C1—C11 | 175.3 (4) | O2—S1—C22—C23 | −126.0 (4) |
N1—C1—C11—C16 | 178.2 (4) | O3—S1—C22—C23 | 3.2 (4) |
N1—C1—C11—C12 | −3.5 (7) | N2—S1—C22—C23 | 120.5 (4) |
C16—C11—C12—O1 | 180.0 (5) | C21—C22—C23—C24 | −0.1 (7) |
C1—C11—C12—O1 | 1.6 (7) | S1—C22—C23—C24 | −179.0 (4) |
C16—C11—C12—C13 | 1.2 (7) | C22—C23—C24—C25 | 0.9 (8) |
C1—C11—C12—C13 | −177.1 (5) | C23—C24—C25—C26 | −0.9 (8) |
O1—C12—C13—C14 | 179.7 (6) | C24—C25—C26—C21 | 0.2 (8) |
C11—C12—C13—C14 | −1.4 (9) | C22—C21—C26—C25 | 0.6 (7) |
C12—C13—C14—C15 | 0.8 (10) | N1—C21—C26—C25 | −176.9 (4) |
C13—C14—C15—C16 | 0.2 (9) | C35—N2—C31—C32 | −55.6 (7) |
C13—C14—C15—Br1 | 179.7 (5) | S1—N2—C31—C32 | 158.4 (5) |
C14—C15—C16—C11 | −0.4 (8) | N2—C31—C32—C33 | 56.1 (8) |
Br1—C15—C16—C11 | −179.9 (4) | C31—C32—C33—C34 | −56.3 (9) |
C12—C11—C16—C15 | −0.2 (7) | C32—C33—C34—C35 | 54.6 (9) |
C1—C11—C16—C15 | 178.1 (4) | C31—N2—C35—C34 | 54.9 (7) |
C1—N1—C21—C22 | 135.0 (5) | S1—N2—C35—C34 | −160.7 (5) |
C1—N1—C21—C26 | −47.6 (6) | C33—C34—C35—N2 | −53.6 (9) |
C26—C21—C22—C23 | −0.6 (7) |
D—H···A | D—H | H···A | D···A | 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 |
Symmetry codes: (i) x+1/2, y, −z+1/2; (ii) x+1/2, −y+1/2, −z. |
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 |
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. |
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 |
Druglikeness | ||
Toxicity | ||
Target | Prediction | Probability |
Hepatotoxicity | Inactive | 0.57 |
Carcinogenicity | Inactive | 0.57 |
Immunotoxicity | Inactive | 0.87 |
Mutagenicity | Inactive | 0.71 |
Cytotoxicity | Inactive | 0.73 |
Druglikeness | ||
Toxicity | ||
Target | Prediction | Probability |
Hepatotoxicity | Inactive | 0.62 |
Carcinogenicity | Inactive | 0.57 |
Immunotoxicity | Inactive | 0.94 |
Mutagenicity | Active | 0.68 |
Cytotoxicity | Inactive | 0.78 |
Druglikeness | ||
Toxicity | ||
Target | Prediction | Probability |
Hepatotoxicity | Inactive | 0.76 |
Carcinogenicity | Inactive | 0.60 |
Immunotoxicity | Inactive | 0.99 |
Mutagenicity | Active | 0.64 |
Cytotoxicity | Inactive | 0.88 |
Druglikeness | ||
Toxicity | ||
Target | Prediction | Probability |
Hepatotoxicity | Inactive | 0.78 |
Carcinogenicity | Active | 0.85 |
Immunotoxicity | Inactive | 0.99 |
Mutagenicity | Inactive | 0.88 |
Cytotoxicity | Inactive | 0.93 |
Compound | Mr | No. of heavy atoms | Fraction CCp3 | 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 | 0.1 | 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 | 0.6 | 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 |
Druglikeness 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 |
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
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. CrossRef Google Scholar
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. Google Scholar
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. Google Scholar
Afsan, Z., Roisnel, T., Tabassum, S. & Arjmand, F. (2020). Bioorg. Chem. 94, 103427. CrossRef PubMed Google Scholar
Akman, F. (2019). Cellul. Chem. Technol. 53, 243–250. CrossRef CAS Google Scholar
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. CrossRef PubMed Google Scholar
Altürk, S., Tamer, Ö., Avcı, D. & Atalay, Y. (2015). J. Organomet. Chem. 797, 110–119. Google Scholar
Becke, A. D. (1993). J. Chem. Phys. 98, 5648–5652. CrossRef CAS Web of Science Google Scholar
Behrens, J. (2000). Ann. N. Y. Acad. Sci. 910, 21–35. CrossRef PubMed CAS Google Scholar
Bruker (2016). APEX2, SAINT, SADABS and TWINABS. Bruker AXS Inc., Madison, Wisconsin, USA. Google Scholar
Choudhary, N., Bee, S., Gupta, A. & Tandon, P. (2013). Comput. Theor. Chem. 1016, 8–21. CrossRef CAS Google Scholar
Dassault Systémes (2020). Discovery Studio. BIOVIA, San Diego, CA, USA. https://www.3ds.com/. Google Scholar
Dueke-Eze, C. U., Fasina, T. M., Oluwalana, A. E., Familoni, O. B., Mphalele, J. M. & Onubuogu, C. (2020). Sci. Afr. 9, e00522. Google Scholar
Eisemann, T. & Pascal, J. M. (2020). Cell. Mol. Life Sci. 77, 19–33. CrossRef CAS PubMed Google Scholar
Elsamra, R. M., Masoud, M. S. & Ramadan, A. M. (2022). Sci. Rep. 12, 20192. CrossRef PubMed Google Scholar
Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854. Web of Science CrossRef CAS IUCr Journals Google Scholar
Feng, X. & Koh, D. W. (2013). Int. Rev. Cell Mol. Biol. 304, 227–281. CrossRef CAS PubMed Google Scholar
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. Web of Science CrossRef PubMed CAS Google Scholar
Gao, C., Xiao, G. & Hu, J. (2014). Cell Biosci. 4, 13. CrossRef PubMed Google Scholar
Gordon, A. T., Hosten, E. C. & Ogunlaja, A. S. (2022). Pharmaceuticals, 15, 1240. CrossRef PubMed Google Scholar
Hall, M. L., Goldfeld, D. A., Bochevarov, A. D. & Friesner, R. A. (2009). J. Chem. Theory Comput. 5, 2996–3009. CrossRef CAS PubMed Google Scholar
Huang, H. & He, X. (2008). Curr. Opin. Cell Biol. 20, 119–125. CrossRef PubMed CAS Google Scholar
Hübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281–1284. Web of Science CrossRef IUCr Journals Google Scholar
Irfan, 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
Jędrzejczyk, M., Janczak, J. & Huczyński, A. (2022). J. Mol. Struct. 1263, 133129. Google Scholar
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. CrossRef IUCr Journals Google Scholar
Kolade, 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
Kondapuram, 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
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. CrossRef CAS PubMed Google Scholar
Kumar, A., Singh, A. K., Singh, H., Vijayan, V., Kumar, D., Naik, J. & Kumar, P. (2023b). Pharmaceuticals, 16, 299. CrossRef PubMed Google Scholar
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. Web of Science CrossRef CAS IUCr Journals Google Scholar
Mahmood, A., Irfan, A. & Wang, J. L. (2022). Chem. Eur. J. 28, e202103712. CrossRef PubMed Google Scholar
Matela, G. (2020). Anticancer Agents Med. Chem. 20, 1908–1917. CrossRef CAS PubMed Google Scholar
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. Google Scholar
Muhammad-Ali, M. A., Jasim, E. Q. & Al-Saadoon, A. H. (2023). J. Med. Chem. Sci. 6, 2128–2139. CAS Google Scholar
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. CrossRef PubMed Google Scholar
Pandi, S., Kulanthaivel, L., Subbaraj, G. K., Rajaram, S. & Subramanian, S. (2022). BioMed Res. Int. 2022, 3338549. CrossRef PubMed Google Scholar
Pawar, S. S. & Rohane, S. H. (2021). Asian J. Res. Chem. 14, 86-–88. CrossRef Google Scholar
Pereira, 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
Prokopenko, Y. S., Perekhoda, L. O. & Georgiyants, V. A. (2019). J. Appl. Pharm. Sci. 9, 066–072. CAS Google Scholar
Salehi, M., Kubicki, M., Galini, M., Jafari, M. & Malekshah, R. E. (2019). J. Mol. Struct. 1180, 595–602. CrossRef CAS Google Scholar
Sheldrick, G. M. (2015a). Acta Cryst. A71, 3–8. Web of Science CrossRef IUCr Journals Google Scholar
Sheldrick, G. M. (2015b). Acta Cryst. C71, 3–8. Web of Science CrossRef IUCr Journals Google Scholar
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. CrossRef CAS PubMed Google Scholar
Silva, R. R. da, Ramalho, T. C., Santos, J. M. & Figueroa-Villar, J. D. (2006). J. Phys. Chem. A, 110, 1031–1040. PubMed Google Scholar
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. Web of Science CrossRef CAS IUCr Journals Google Scholar
Spek, A. L. (2020). Acta Cryst. E76, 1–11. Web of Science CrossRef IUCr Journals Google Scholar
Trott, O. & Olson, A. J. (2010). J. Comput. Chem. 31, 455–461. Web of Science CrossRef PubMed CAS Google Scholar
Yele, 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.