Abstract
The escalating threat of antimicrobial resistance calls for novel therapeutic agents. This study employed a ligand-based design approach to develop three series of N-arylpyrrole derivatives (Va–e, VIa–e, and VIIa–e), refined through molecular modeling. Synthesized compounds were evaluated against ESKAPE pathogens, MRSA, and Mycobacterium phlei. Series Va–e showed the most promise, with compounds Vb, Vc, and Ve outperforming levofloxacin against MRSA (MIC = 4 μg/mL vs. 8 μg/mL). Vc also exhibited activity against E. coli, K. pneumoniae, and A. baumannii, and showed significant inhibition against M. phlei (MIC = 8 μg/mL). Compounds were evaluated for antibiofilm and antivirulence properties, targeting resistance mechanisms linked to infection persistence and dissemination. Most exhibited broad-spectrum biofilm inhibition and antivirulence activity. Cytotoxicity studies revealed selectivity for bacterial cells. ADMET studies supported drug-like properties. Docking studies suggested UPPP inhibition as the potential mechanism. SAR analysis was conducted to support future optimizations.
Keywords: N-arylpyrrole, antibacterial, anti-virulence, antitubercular, ESKAPE
Graphical abstract
Introduction
Antimicrobial resistance poses a significant threat to global health, as antibiotics are indispensable for basic and modern medical practices. Cancer chemotherapy, invasive surgeries, organ transplants, and complicated deliveries may only be performed without the danger of deadly infections if appropriate anti-bacterial medicines are available1. Projections suggest a staggering 10 million deaths attributable to infectious disease treatment failure by 20502.
Among the most concerning are the ESKAPE pathogens, a group comprising Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter species. These bacteria, collectively dubbed “ESKAPE” due to their remarkable ability to escape the effects of antibiotics, have become emblematic of the challenges posed by antimicrobial resistance3. Of particular concern within this group is methicillin-resistant Staphylococcus aureus (MRSA), which exemplifies the significant resistance seen in certain bacterial strains. MRSA has developed resistance to various antibiotics, including beta-lactams, aminoglycosides, glycopeptides, and fluoroquinolones4. The global prevalence of MRSA is alarming, with numerous reports indicating a substantial proportion of Staphylococcus infections resistant to methicillin5,6.
Furthermore, gram-negative pathogens such as Pseudomonas aeruginosa, Acinetobacter baumannii, Klebsiella pneumoniae, and Escherichia coli are increasingly prevalent and demonstrate resistance to conventional therapy. Pseudomonas aeruginosa, in particular, poses a global threat due to its rapid development of resistance during treatment, often facilitated by forming biofilm aggregates7–9. Similarly, certain strains of Acinetobacter baumannii exhibit multidrug resistance, potentially attributable to their adeptness in deploying diverse resistance mechanisms10. Epidemic strains of K. pneumoniae have emerged resistant to carbapenem antibiotics11 and fluoroquinolones12. Moreover, infections caused by Escherichia coli, particularly urinary tract infections, are increasingly resistant to third-generation cephalosporins in certain regions, complicating treatment13.
To confront the swiftly escalating crisis of resistance, it is imperative to undertake extensive research and develop innovative antimicrobial agents. This research aims to contribute to global efforts in combating antimicrobial resistance and developing effective therapies for infectious diseases.
A literature review exploring effective chemical moieties with potent anti-bacterial activity, particularly against resistant strains, has identified pyrroles as a promising antimicrobial scaffold. Various pyrrole derivatives have been synthesised and evaluated for their antimicrobial properties. For example, tetra-substituted pyrrole derivatives were synthesised and tested against Escherichia coli and S. aureus (1, Figure 1)14. Furthermore, diphenyl pyrrole compounds exhibited notable activity against drug-resistant gram-positive and gram-negative bacteria, comparable to or surpassing levofloxacin (2, Figure 1)15. Notably, pyrrole derivatives have shown significant anti-tubercular activity as well. For instance, BM212, a 2,5-dimethylpyrrole derivative currently undergoing clinical phase II trials, demonstrates potent activity against various mycobacteria, including multidrug-resistant (MDR) tuberculosis strains (3, Figure 1)16.
Figure 1.
Tetra-substituted pyrrole-based derivatives (1), 1,5-diphenyl pyrrole derivatives (2), antitubercular drug BM212 (3), phenylthiazole lead compound (4), phenylpyrazole derivatives (5), and (thio)semicarbazide derivatives (6,7).
Apart from pyrroles, a highly active phenylthiazole scaffold with significant anti-bacterial effects against gram-positive bacteria, including 18 MRSA and VRE strains, was identified.17. Structure–activity relationship (SAR) analysis revealed a thiazole nucleus attached to two key structural components: a highly lipophilic moiety at C-2, corresponding to greater potency, and a cationic moiety at C-5 essential for anti-bacterial activity (4, Figure 1)18. Despite its anti-MRSA activity, this scaffold faced pharmacokinetic limitations17,19. The central thiazole ring of this scaffold was replaced with a pyrazole ring while retaining the other pharmacophoric features, resulting in compounds with notable activity against gram-positive and gram-negative bacteria (5, Figure 1)17.
Furthermore, our literature review revealed the bio-potential of semicarbazides and thiosemicarbazide derivatives as non-toxic anti-bacterial agents (6,7) (Figure 1)20–23. Incorporating these derivatives into our design may yield active anti-bacterial agents.
In this work, we hypothesised that incorporating a dimethylpyrrole ring while maintaining the other pharmacophoric features of the phenylthiazole lead could result in compounds with a broad spectrum (Figure 2). Additionally, we utilised semicarbazide and thiosemicarbazide for the cationic part of our scaffold, aiming to potentially enhance the interaction of the compounds with bacterial targets and improve anti-bacterial potency.
Figure 2.
Design strategy.
A ligand-based preliminary evaluation of the designed compounds using 3DQSAR modelling and ligand pharmacophore mapping protocols was done to predict the potential activity of the newly designed scaffold against MRSA. Subsequently, the designed compounds were synthesised, purified, and structurally confirmed using various analytical and spectral techniques.
The anti-bacterial activities of the synthesised compounds were assessed against a panel of bacteria, including ESKAPE pathogens, including MRSA, and anti-tubercular activity against Mycobacterium phlei. Their antivirulence properties were also assessed using biofilm inhibition and virulence factor assays. Furthermore, the cytotoxicity of the four most potent compounds (Vb, Vc, Vd, and Ve) was assessed.
Additionally, in silico target prediction and docking studies were performed to gain preliminary insights into the potential mechanism of action. ADMET studies were conducted to evaluate the synthesised derivatives’ physicochemical, pharmacokinetic, and toxicity parameters. Finally, a SAR study was established to guide future structural modifications.
Results and discussion
Preliminary evaluation of the designed compounds using computer-aided molecular modelling
Ligand-based computer-aided drug design is used to design and evaluate new drug leads by employing ligands with known activity to help develop new biologically active compounds, allowing us to make better-informed decisions, hence exhausting fewer resources24,25.
This work used consecutive computer-aided drug design protocols to preliminary evaluate our newly designed compounds. First, we generated a ligand-based 3D QSAR pharmacophore using a dataset of phenylthiazole and phenylpyrazole compounds with known activity (MIC) against MRSA26–28. The HypoGen algorithm summarised the structural features of these compounds to generate a valid predictive pharmacophore model using the Discovery Studio 2024 software package29. Then, the phenylthiazole lead used to design our molecules17 was mapped on the generated pharmacophore model (Hypo1) via the ligand pharmacophore mapping protocol for validation. Next, the designed compounds were mapped on the validated pharmacophore model Hypo1 to predict their activity.
3D QSAR pharmacophore model generation
3D QSAR (Three-Dimensional Quantitative Structure–Activity Relationship) pharmacophore model was generated using DS by utilising a dataset of 32 phenylthiazole and phenylpyrazole compounds with known activity (MIC) against MRSA that were extracted from literature26–28. The data set was randomly divided into 22 compounds as a training set (Figure 3) and 10 as a test set (Figure 4). The HypoGen algorithm generated pharmacophore models with the essential features responsible for the anti-MRSA activity. Ten hypotheses were generated using the HypoGen algorithm and validated for their statistical significance and accuracy of prediction (Table 1). After careful review, Hypo1 was selected as the most valid based on the cost difference, root mean square (RMS) difference, and correlation coefficient hypothesis and used for this study.
Figure 3.
Training set ligands along with their MIC values (μg/mL).
Figure 4.
Test set ligands along with their MIC values (μg/mL).
Table 1.
Statistical parameters of the top 10 generated pharmacophore models.
| Hypo no. | Maximum fit | Total cost | Cost distance | RMS | Correlation coefficient (r) | Features |
|---|---|---|---|---|---|---|
| 1 | 9.89 | 96.60 | 102.71 | 1.58 | 0.90 | HBA, HYD, HYD, HYD, RA |
| 2 | 8.66 | 101.52 | 97.797 | 1.70 | 0.88 | HBA, HYD, HYD, HYD, RA |
| 3 | 7.35 | 101.78 | 97.444 | 1.68 | 0.88 | HYD, HYD, HYD, PI, RA |
| 4 | 6.03 | 123.06 | 76.305 | 2.18 | 0.80 | HBA, HYD, HYD, RA |
| 5 | 6.05 | 123.15 | 76.165 | 2.19 | 0.80 | HYD, HYD, HYD, PI, RA |
| 6 | 6.25 | 125.44 | 73.878 | 2.22 | 0.79 | HBA, HYD, HYD, HYD, RA |
| 7 | 7.91 | 125.46 | 73.858 | 2.22 | 0.82 | HBA, HYD, HYD, HYD, RA |
| 8 | 5.86 | 127.76 | 72.174 | 2.24 | 0.78 | HBA, HBA, HYD, HYD |
| 9 | 4.92 | 128.63 | 70.683 | 2.31 | 0.77 | HBA, HYD, RA |
| 10 | 5.66 | 129.67 | 69.642 | 2.26 | 0.78 | HBA, HYD, HYD, HYD, RA |
Validation of selected pharmacophore model
The Discovery Studio software employed the HypoGen algorithm to compute three validation cost functions. The first, termed the fixed cost, assumes a straightforward hypothesis model that smoothly accommodates all dataset and library molecules, resulting in the lowest cost30. The fixed cost value for the generated hypothesis is equal to 69.07. Conversely, the null cost represents the highest potential error cost31. The null cost for the Hypotheses generated is equal to 199.32. Each pharmacophore hypothesis undergoes an independent calculation of total cost, comprising weight, error, and fixed costs32. Among the 10 generated hypotheses, total costs ranged from 96.6 to 129.67. Hypo1 displayed the most significant cost difference of 102.71 (Table 1), indicating that the pharmacophore model is more than 90% statistically significant33. Furthermore, Hypo1 also scored the highest correlation coefficient value of 0.90 (Table 1S) and the lowest RMS value of 1.58, both attributed to a superior capacity for biological activity prediction33.
Furthermore, the predictive ability of the pharmacophore models was assessed by mapping a test set of 10 diverse compounds with known MIC values against MRSA using the Ligand Pharmacophore Mapping tool. A strong correlation (R2 = 0.76) between estimated and reported activity confirmed the model’s reliability (Table 2S). The selected pharmacophore model (Hypo1) (Figure 5) consists of five key features: one hydrogen bond acceptor (HBA), three hydrophobic (HYD), and one ring aromatic (RA). Hypo1 was subsequently used to evaluate the novel scaffold.
Figure 5.
Spatial arrangement of the valid pharmacophore model (Hypo1) with the distances and angles displayed.
Mapping of the lead compound on the valid generated model
The phenylthiazole lead compound was also mapped via the Ligand Pharmacophore Mapping protocol in DS on the chosen pharmacophore model (Hypo1). The lead compound successfully mapped into the pharmacophore’s five features (Figure 6) with a high fit value of 7.8 and estimated activity of 7 µg/mL, further validating Hypo1.
Figure 6.
The mapping of the phenylthiazole lead into the generated pharmacophore model (Hypo1).
Ligand pharmacophore mapping
To preliminary evaluate our newly designed scaffold, the proposed compounds were mapped into the generated pharmacophore using the Ligand Pharmacophore Mapping protocol. All the newly designed compounds were mapped successfully into at least four features of the valid generated pharmacophore model, indicating potential anti-bacterial activity (Table 3S). The fit values ranged between 3.3 and 6.8. We noticed that elongating the compound’s lipophilic tail (R2) allowed it to fit into all five pharmacophore features. This note was incorporated into our design as it might indicate higher activity (Figure 7).
Figure 7.
The mapping of compounds Va and Vc into the generated pharmacophore model.
Chemistry
Scheme 1 includes the synthesis of the intermediates IVa–e. First, the Paal–Knorr condensation reaction of 2,5-hexanedione (I) and different anilines IIa–c in the presence of sulphamic acid as a catalyst yielded the N-aryl-2,5-dimethylpyrrole derivatives IIIa–c34. Second, the N-aryl-2,5-dimethyl-1H-pyrrole-3-carbaldehydes IVa–c and the 2,5-dimethyl-1-aryl-1H-pyrrol-3-yl-ethanones IVd–e were obtained via the simple and convenient Vilsmeier–Haack method in good yields35,36. All the products were identified using TLC and melting points that matched the reported ones37–43. The intermediates IVa–e were also confirmed by 1H NMR and IR spectroscopy, which was in accordance with the literature39,44–46. The 1H NMR spectrum of the novel intermediate IVc showed two singlet signals at δ 1.93 and 2.29 ppm corresponding to the methyl groups at position five and two on the pyrrole ring, respectively. The 1H NMR spectrum also showed one singlet at δ 6.3 ppm corresponding to the pyrrole aromatic proton and an aldehydic proton at δ 9.88 ppm. In addition, the IR spectra for the compound showed signal at 1730 cm−1 for the C═O group.
Scheme 1.
Reagents and conditions: (1) sulphamic acid, rt, 1–6 h; (2) POCl3, ice cold DMF, 60, 4 h, 20% NaOH; (3) POCl3, ice cold DMA, reflux, 6 h, 20% NaOH.
Scheme 2 shows the reaction of the intermediates IVa–e with aminoguanidine carbonate, thiosemicarbazide HCl, and semicarbazide HCl, thus replacing the acetyl/formyl moiety in the intermediates IVa–e with various cationic tails, such as aminoguanidine, semicarbazide, and thiosemicarbazide, affording the desired final compounds Va–e, VIa–e, and VIIa–e, respectively. The reaction is based on nucleophilic addition on the carbonyl group of the appropriate aldehyde or ketone; the partial positive charge on the carbonyl atom makes it susceptible to attack by a nucleophile like a primary amine forming imine derivatives, also known as Schiff bases (compounds having a C═N function)47. The structures of compounds Va–e, VIa–e, and VIIa–e were supported by elemental analyses, high-resolution mass spectrometry (HRMS), and spectroscopic data: IR, 1H NMR, and 13C NMR.
Scheme 2.
Reagents and conditions: (4) aminoguanidine HCO3, conc. HCl, absolute ethanol, reflux, 20–24 h; (5) thiosemicarbazide HCl, anhydrous sodium acetate, absolute ethanol, reflux, 12 h; (6) semicarbazide HCl, anhydrous sodium acetate, absolute ethanol, reflux, 12 h.
The IR spectra of all series compounds showed IR signals at 3289–3565 cm−1 corresponding to the NH/NH2 groups. Compounds Va–e showed IR signals at 1582–1670 cm−1 corresponding to the C═N group in the aminoguanidine moiety, while compounds VIa–e showed IR signals at 1482–1521 cm−1 corresponding to the C═S group in the thiosemicarbazide moiety, and compounds VIIa–e showed IR signals at 1681–1687 cm−1 corresponding to the C═O group in the semicarbazide moiety.
The 1H NMR spectra of compounds VIa–e showed new exchangeable singlet signals at δ 5.51 and 5.8 ppm, corresponding to the NH and NH2 groups. Compounds VIa–e and VIIa–e displayed one singlet at δ 6.17 ppm corresponding to the NH2 moiety and another at δ 9.02–9.72 ppm corresponding to the NH group. Furthermore, 1H NMR for compounds Va–c, VIa–c, and VIIa–c showed the disappearance of the aldehydic proton at δ 9.88 ppm and the appearance of a singlet at δ 7.99–8.75 ppm corresponding to the proton at the CH═N. While compounds Vd–e, VId–e, and VIId–e showed an extra singlet at δ 2.16–2.3 ppm corresponding to the methyl group at R2.
The Va–e, VIa–e, and VIIa–e structures were further assured by 13C NMR spectra, which shared the same essential peaks at δ 140.8–158 ppm corresponding to the C═NH moiety. Compounds Va–e showed a peak at δ 158.3–158.7 ppm corresponding to the C═N group in the aminoguanidine moiety, while compounds VIa–e showed peaks at δ 194.2–199.2 ppm corresponding to the C═S group in the thiosemicarbazide moiety, and compounds VIIa–e showed peaks at δ 158 ppm corresponding to the C═O group in the semicarbazide moiety. Compounds Vd–e, VId–e, and VIId–e showed an extra signal at δ 16.1–16.4 ppm corresponding to the methyl group at R2.
Biological evaluation
Antimicrobial activity
All the synthesised compounds were tested for their biological activity by determining their MIC values against a panel of ESKAPE pathogens, including MRSA, and anti-tubercular activity against Mycobacterium phlei using the broth microdilution method. In addition, all compounds were tested for their anti-tubercular activity against clinical isolates of Mycobacterium phlei. The anti-bacterial drugs levofloxacin and rifampicin were used as reference standards. Compounds displaying MIC values lower or equal to the reference standard drug were considered promising. The results of the antimicrobial evaluation are presented in Table 2.
Table 2.
The minimum inhibitory concentration (MIC in μg/mL) of the all the synthesised compounds against a panel of susceptible and drug-resistant gram + ve, gram − ve bacteria and clinical isolates of Mycobacterium phlei.
|
Gram-positive strains (µg/mL) | Acid-fast strains (µg/mL) | Gram-negative strains (µg/mL) |
|||||||
|---|---|---|---|---|---|---|---|---|---|---|
| Enteric | Non-enteric | |||||||||
| Compound | R1 | R2 | x |
MSSA ATCC 29213 |
MRSA Clinical isolate |
Mycobacterium phlei Clinical isolate |
E. coli ATCC 25922 |
K. pneumoniae ATCC 700603 |
A. baumannii Clinical isolate |
P. aeruginosa ATCC 27853 |
| Va | H | H | NH | 16 | 32 | 128 | 128 | 128 | 128 | 256 |
| Vb | Cl | H | NH | 4 | 4 | 16 | 8 | 8 | 16 | 16 |
| Vc | Ph | H | NH | 4 | 4 | 8 | 4 | 8 | 8 | 32 |
| Vd | H | CH3 | NH | 16 | 8 | 64 | 32 | 64 | 32 | 128 |
| Ve | Cl | CH3 | NH | 8 | 4 | 32 | 32 | 16 | 32 | 128 |
| VIa | H | H | S | >1024 | 1024 | 1024 | >1024 | 1024 | 1024 | 1024 |
| VIb | Cl | H | S | 256 | 128 | 128 | 512 | 256 | 128 | 1024 |
| VIc | Ph | H | S | 64 | 64 | 128 | 512 | 512 | 64 | 128 |
| VId | H | CH3 | S | 1024 | 1024 | 512 | 1024 | 1024 | 256 | 1024 |
| VIe | Cl | CH3 | S | 64 | 64 | 128 | 256 | 512 | 256 | 256 |
| VIIa | H | H | O | >1024 | 512 | 128 | >1024 | >1024 | 512 | 1024 |
| VIIb | Cl | H | O | 256 | 256 | 256 | 512 | 512 | 256 | 128 |
| VIIc | Ph | H | O | 256 | 64 | 128 | 128 | 256 | 256 | 256 |
| VIId | H | CH3 | O | >1024 | 1024 | 512 | >1024 | 1024 | 1024 | 1024 |
| VIIe | Cl | CH3 | O | 64 | 64 | 128 | 1024 | 512 | 256 | 256 |
| Levofloxacin | 0.5 | 8 | 2 | 0.05 | 4 | 8 | 4 | |||
| Rifampicin | 0.015 | 4 | 0.0125 | 16 | 8 | 8 | 64 | |||
Anti-bacterial activity against gram + ve strains
In the evaluation of anti-bacterial activity against both methicillin-susceptible S. aureus (MSSA) and clinical isolates of MRSA (Table 2), the aminoguanidine series Va–e demonstrated excellent activity against both MSSA and MRSA with MIC values ranging from 4 μg/mL to 32 μg/mL. Compounds Vb, Vc, and Ve outperformed levofloxacin (MIC = 8 μg/mL) with a MIC of 4 μg/mL on MRSA, revealing promising activity as anti-MRSA agents. Vd showed notable activity against MRSA with a MIC of 8 μg/mL, equivalent to levofloxacin. Compound Va displayed the least activity in the series against MRSA with an MIC of 32 μg/mL. In the thiosemicarbazide series, compounds VIc and VIe displayed a MIC value of 64 μg/mL for both MSSA and MRSA, while compounds VIa, VIb, and VId showed higher MIC values (≥512 μg/mL). Similarly, in the semicarbazide series, compounds VIIc and VIIe exhibited a MIC value of 64 μg/mL for both MSSA and MRSA, while VIIa, VIIb, and VIId displayed little to no activity against both strains (≥256 μg/mL).
Anti-bacterial activity against gram − ve strains
All synthesised compounds were tested for their anti-bacterial activity against the enteric gram-negative bacterial strains E. coli and K. pneumoniae and the non-enteric gram-negative bacterial strains A. baumannii and P. aeruginosa (Table 2). In the aminoguanidine series Va–e, Vc exhibited notable activity with MIC values of 4 μg/mL against E. coli, 8 μg/mL against K. pneumoniae, and 8 μg/mL against A. baumannii, comparable to levofloxacin. Vb also displayed significant activity with a MIC of 8 μg/mL against both enteric strains and 16 μg/mL against non-enteric strains. The thiosemicarbazide series VIa–e showed less promising results, as VIb and VIe demonstrated higher MIC values of 256 μg/mL against K. pneumoniae and E. coli, respectively, while VIa, VIc, and VId exhibited MIC values ≥512 μg/mL. The semicarbazide series VIIa–e displayed little to no activity against enteric and non-enteric bacterial strains, with MIC values ranging from 128 μg/mL to 1024 μg/mL.
Anti-tubercular activity against Mycobacterium phlei
The anti-tubercular activity of synthesised compounds against clinical isolates of Mycobacterium phlei was assessed (Table 2). In the aminoguanidine series Va–e, Vc exhibited the most potent anti-tubercular effect with the lowest MIC value of 8 μg/mL, followed by Vb with a MIC of 16 μg/mL. Compounds Va, Vd, and Ve displayed MIC values ranging from 32 μg/mL to 128 μg/mL. For the thiosemicarbazide series, VIa–e, VI, VIc, and VIe demonstrated the lowest MIC value of 128 μg/mL, while VIa and VId exhibited higher MIC values of 1024 μg/mL and 512 μg/mL, respectively. The semicarbazide series VIIa–e showed MIC values ranging from 128 μg/mL to 512 μg/mL.
Anti-virulence activity
Bacterial biofilms are organised communities of bacteria that attach to surfaces and become encased within a self-produced extracellular polymeric substance (EPS) matrix. This matrix, composed of polysaccharides, proteins, nucleic acids, and lipids, provides structural integrity, protection, and an environment conducive to bacterial communication and nutrient exchange48 The EPS matrix also acts as a protective barrier, shielding bacteria from environmental stressors such as antibiotics, desiccation, and immune responses49. We evaluated the antibiofilm efficacy of the designed compounds at sub-MIC doses by measuring the biofilm density of treated cultures relative to untreated control. Notably, all the compounds exhibited significant, broad-spectrum anti-biofilm activity against biofilms formed by E. coli, S. aureus, and P. aeruginosa (A, Figure 8). Microscopic analysis of biofilms from E. coli, and S. aureus culture treated with the most potent compounds revealed distinct differences in biofilm structure between untreated and treated bacterial cultures. Untreated biofilms exhibited dense, cohesive structures, while drug-treated biofilms appeared disrupted, sparse, and unstable, reflecting effective antibiofilm activity (B, Figure 8).
Figure 8.
Evaluation of the anti-virulence effects of designed compounds. (A) Antibiofilm activity of designed compounds against E. coli, S. aureus, and P. aeruginosa. Biofilm density was quantified and expressed as a percentage relative to untreated controls. Statistical analysis was performed using one-way ANOVA followed by Dunnett’s multiple comparison test (***p < 0.001). (B) Microscopic visualisation of biofilms from E. coli and S. aureus. Scale bar 50 µm. (C) Anti-protease activity of designed compounds evaluated in S. aureus and P. aeruginosa using skimmed milk agar. Protease activity in untreated cultures produced clear inhibition zones, which were entirely absent in drug-treated cultures, confirming complete protease inhibition. U: untreated culture; S: 1% SDS; N: no bacterial lysate.
Proteases represent another critical bacterial virulence factor, facilitating infection by breaking down host tissues and evading immune defences50. To evaluate the impact of the designed compounds on protease activity, cultures of P. aeruginosa and S. aureus were treated with the drugs and analysed using the skimmed milk agar method. Supernatants from untreated cultures served as controls, demonstrating full protease activity (C, Figure 8). Protease activity was evidenced by clear inhibition zones surrounding the wells, with larger zones corresponding to higher enzymatic activity. In cultures treated with the compounds, these inhibition zones were completely absent, indicating effective protease inhibition. Collectively, these findings highlight the remarkable ant virulence properties of the designed compounds.
In vitro cytotoxicity assay
Antimicrobial agents should demonstrate low or no toxicity towards mammalian cells51,52. Ideally, a potential drug should have a relatively high cytotoxic concentration (CC50) and a relatively low active concentration53. Evaluation of the selectivity index (SI) value is considered crucial for determining the therapeutic potential of evaluated compounds54. Selectivity index is defined as the ratio of the toxic concentration known as 50% cell cytotoxic concentration (CC50) of a sample against its effective bioactive concentration (MIC)55. High values of SI indicate excellent selectivity55–57.
This work evaluated the cytotoxicity of the four most potent compounds (Vb, Vc, Vd, and Ve) by MTT assay against normal mammalian cell line VERO (African Green Monkey Kidney cells). The CC50 values were calculated for the four compounds, along with the SI of each compound towards MRSA (Table 3). In this work, SI was calculated following this equation:
where CC50 is 50% vero cell cytotoxic concentration for 48 h in µg/mL, and MIC is minimal inhibitory concentration against MRSA in µg/mL.
Table 3.
In vitro cytotoxicity screening against normal mammalian cell line VERO for the most potent compounds, their MIC on MRSA, and calculated mean selectivity index.
| Cpd | Cytotoxicity | Active Concentration | Mean SI |
|---|---|---|---|
| Vero CC50 (µg/mL) | MRSA MIC value (µg/mL) | ||
| Vb | 43.32 ± 2.86 | 4 | 10.83 |
| Vc | 188.17 ± 6.87 | 4 | 47.01 |
| Vd | 2.8 ± 0.35 | 8 | 0.27 |
| Ve | 10.33 ± 0.49 | 4 | 2.55 |
The most potent compound, Vc, demonstrated excellent selectivity towards MRSA with a mean SI of 47.01 and CC50 value of 188.17 ± 6.87 µg/mL; moreover, cell viability was 100% at a concentration of 15.6 µg/mL, which is almost four times its MIC against MRSA (Table 3). Compound Vb also demonstrated good selectivity against MRSA with a mean SI value of 10.83 and CC50 value of 43.32 ± 2.86 µg/mL. Moreover, cell viability was 76.38% at a concentration of 15.6, almost four times its MIC against MRSA (Figure 1S). Compounds Vd and Ve displayed moderate to weak selectivity against MRSA with mean SI values of 0.27, 2.55, and CC50 values of 2.8 ± 0.35 and 10.33 ± 0.49 µg/mL, respectively.
In silico studies
Docking studies and target identification
Bacterial cytological profiling (BCP) previously indicated that the phenylthiazole lead compound exerted rapid bactericidal activity via inhibition of cell wall synthesis, resulting in cell lysis and morphological defects19,58. Further validation using transposon mutagenesis pinpointed undecaprenyl pyrophosphate phosphatase (UPPP) as the primary target19.
UPPP plays a crucial role in bacterial isoprenoid biosynthesis, an early and essential step in peptidoglycan formation59. Unlike traditional antibiotics such as β-lactams and glycopeptides that interfere with late-stage transpeptidation or crosslinking, UPPP inhibition disrupts the initial stages of cell wall biosynthesis – a mechanism particularly advantageous against resistant strains like MRSA and VRE60–63
To explore whether our newly designed N-arylpyrrole derivatives act through a similar mechanism, we docked them into the UPPP active site (PDB: 6CB2) using the CDOCKER protocol in Discovery Studio 202464. The phenylthiazole lead showed a -CDOCKER energy of 22.72 and formed four key hydrogen bonds – three involving the aminoguanidine group and Glu49/Glu17, and one involving the thiazole sulphur and Thr28 (Figure 9).
Figure 9.
The 2D interactions of the lead compound and our designed compound Vc in the binding site of UPPP protein (PDB code: 6CB2).
Encouragingly, our designed arylpyrrole derivatives exhibited comparable docking poses and favourable binding energies, suggesting potential UPPP inhibition as a shared mechanism of action (Figure 9, Table 4S). Despite structural differences, the conserved binding interactions imply that these compounds may disrupt bacterial cell wall synthesis by targeting the same pathway.
We therefore hypothesise that the observed broad-spectrum antibacterial, antibiofilm, and antivirulence activities of the N-arylpyrrole series are at least partly mediated through inhibition of UPPP, leading to impaired peptidoglycan biosynthesis and cell wall integrity. This proposed mechanism aligns with the potent activity against MRSA and ESKAPE pathogens and justifies further studies to validate UPPP as a primary molecular target.
ADMET studies
Absorption, distribution, metabolism, excretion, and toxicity (ADMET) are essential parameters to be considered during the drug development process, as they provide a base for the newly designed compounds to be efficient drug candidates65,66. Therefore, in silico ADMET studies were performed by applying the ADMET descriptors algorithm in Discovery Studio 2024 software package67. In which pharmacokinetic parameters such as aqueous solubility68,69, blood–brain barrier (BBB) penetration70, cytochrome P4502D6 (CYP2D6) inhibition71, human intestinal absorption (HIA)72, plasma protein binding (PPB)73 were calculated quantitatively for the four most potent compounds (Vb, Vc, Vd, and Ve) (Table 4).
Table 4.
In silico ADMET screening of the synthesised compounds.
| Compound | Solubility level | BBB-level | Absorption level | ALog P98 | PSA_2D | CYP2D6# prediction | PPB # prediction |
|---|---|---|---|---|---|---|---|
| Vb | 3 | 3 | 0 | 2.69 | 79.109 | True | True |
| Vc | 2 | 2 | 0 | 3.54 | 79.109 | False | True |
| Vd | 3 | 3 | 0 | 2.01 | 79.109 | False | True |
| Ve | 3 | 3 | 0 | 2.67 | 79.109 | False | True |
Comparisons were made with reference level values obtained from Accelrys Discovery Studio (Table 4S)74,75. Compounds Vb, Vd, and Ve, scoring at level 3, showed low BBB penetration, suggesting a low risk of central nervous system (CNS) side effects. In contrast, compound Vc, scoring at prediction level 2, indicated medium BBB penetration70. None of the evaluated compounds, except for Vb, inhibited CYP2D6, thus minimising the likelihood of drug–drug interactions71. They were also predicted to have high PPB; however, modifications of these compounds could potentially enhance their drug-like properties.
Polar surface area (PSA) plays a significant role in HIA and membrane permeability76. Both PSA and AlogP98 were calculated to assess drug bioavailability. Molecules with PSA > 150 or AlogP98 > 7.0 are typically associated with very low absorption and bioavailability77. All evaluated compounds exhibited a PSA of 79.10 and AlogP98 values ranging from 2.01 to 3.54, indicating good oral absorption theoretically78. The intestinal absorption model, depicted in Figure 10, incorporates 95% and 99% confidence ellipses in the ADMET_PSA_2D, ADMET_AlogP98 plane. These ellipses delineate regions where well-absorbed compounds are anticipated: 95% of well-absorbed compounds are expected within the 95% ellipse, while 99% should fall within the 99% ellipse. All evaluated compounds fell within the 99% ellipse, indicating good absorption.
Figure 10.
Plot of human intestinal absorption (HIA) and blood–brain barrier plot for the newly synthesised hits.
Structure–activity relationship study
A brief SAR study was extracted from the results obtained from the antimicrobial evaluation of the novel N-arylpyrrole derivatives. Judging by the MIC values of the fifteen compounds evaluated, it was found that the aminoguanidine series Va–e were the most active derivatives. In agreement with our preliminary evaluation, compound Vc, featuring a long and highly lipophilic tail (phenyl moiety), stood out as the most potent with promising activity on MRSA (4 μg/mL), E. coli (4 μg/mL), A. baumannii (8 μg/mL), K. pneumoniae (8 μg/mL), and P. aeruginosa (32 μg/mL). Vc also exhibited the lowest MIC value (8 μg/mL) of all tested compounds against clinical isolates of Mycobacterium phlei. Compound Vb, with the second most lipophilic tail (chloride moiety), also displayed promising activity, albeit with slightly higher MIC values. Adding a methyl group at R2 in compound Ve resulted in one-fold higher MIC values, except for the MIC value against MRSA, which remained the same (4 μg/mL). Compound Va, belonging to the aminoguanidine series Va–e and containing no lipophilic tail (R1═H), had the least activity against all tested strains; however, the addition of methyl group at R2 in compound Vd increased the activity, especially against MRSA, where the MIC value of Vd was eightfolds lower than Va (4 μg/mL).
Thiosemicarbazide and semicarbazide series VIa–e and VIIa–e generally exhibited minimal antimicrobial activity. However, compounds VIc, VIIc, VIe, and VIIe, with more lipophilic tails, showed a MIC of 64 μg/mL against MRSA. Increasing lipophilicity at R1 significantly enhanced antimicrobial activity and adding a methyl group at R2 further boosted the activity. In conclusion, aminoguanidine substitution, increased lipophilicity at R1, and adding a methyl group at R2 are crucial factors for achieving broad-spectrum anti-bacterial activity in the investigated arylpyrrole scaffold. A graphical representation of the SAR is presented in Figure 11.
Figure 11.
Graphical representation of the SAR.
Conclusions
In this study, we synthesised and evaluated three series of N-arylpyrrole derivatives (Va–e, VIa–e, and VIIa–e) for their antimicrobial potential. Series Va–e, particularly compounds Vb, Vc, and Ve, exhibited significant anti-bacterial activity against MRSA and other ESKAPE pathogens. Notably, compound Vc demonstrated superior potency with a MIC of 4 μg/mL against MRSA, outperforming levofloxacin (MIC = 8 μg/mL), and showed activity against E. coli, K. pneumoniae, and A. baumannii. Additionally, Vc displayed notable anti-tubercular activity against M. phlei (MIC = 8 μg/mL). Beyond their direct antimicrobial effects, the compounds exhibited significant antivirulence properties, disrupting biofilm formation in E. coli, S. aureus, and P. aeruginosa at sub-MIC levels and effectively inhibiting bacterial protease activity in P. aeruginosa and S. aureus. Cytotoxicity assessments confirmed Vb and Vc as highly selective for bacterial cells, with Vc displaying an impressive SI of 47.01. In silico ADMET predictions supported favourable pharmacokinetic and toxicity profiles for the synthesised compounds, suggesting their potential as lead candidates for further development. Docking studies further predicted that the antibacterial mechanism involves inhibition of the UPPP enzyme, a key player in early-stage bacterial cell wall synthesis. This may explain their activity against resistant strains such as MRSA. These findings position N-arylpyrrole derivatives, particularly Vc, as compelling leads for the development of new antimicrobial agents. Further studies including structural refinement, resistance development profiling, and in vivo validation are warranted to comprehensively evaluate their therapeutic potential.
Experimental
Preliminary evaluation of the designed compounds using computer-aided molecular modelling
3D QSAR pharmacophore generation
The Discovery Studio V4.1 program performed all the molecular modelling studies. All the compounds were sketched in the Discovery studio program V4.1 and optimised using a CHARMM force field. For the 3D QSAR model generation, a library of 32 phenylthiazole and phenylpyrazole derivatives with reported activity (MIC) against MRSA was extracted from the literature26–28 and randomly divided into training and test sets. The training set was subjected to the feature mapping protocol in DS to identify distinct chemical features present in the ligands. The features revealed were HBA, hydrogen bond donor (HBD), HYD, and RA. The four features identified were chosen for the 3D QSAR pharmacophore generation protocol. The MIC was selected as the active property, and the energy threshold was retained at 20 kcal/mol throughout the protocol run. The uncertainty value was set to 1.5. This value represents a ratio of the reported value to the minimum and maximum values. Setting the uncertainty value to 1.5 entails that the model can acclimate differences in the experimental MIC values and predict MIC up to 1.5 times32. All the other parameters were left to default.
The HypoGen algorithm utilised in the 3D QSAR pharmacophore generation protocol of DS interpreted the common chemical features related to low or high biological activity in the training set, and 10 hypotheses were generated. Cost analysis was used to choose the best hypothesis from the 10 generated models.
Chemistry
Chemicals were purchased from Sigma-Aldrich (Darmstadt, Germany), Merck (Darmstadt, Germany), Alfa Aesar (Karlsruhe, Germany), and Loba Chemie (Mumbai, India) and were used as such without further purification. Reactions were followed using analytical thin layer chromatography (TLC), performed on Aluminium silica gel 60 F254 TLC plates purchased from Merck (Darmstadt, Germany), with visualisation under UV light (254 nm). Melting points were recorded on Stuart Scientific apparatus and are reported herein uncorrected. Routine 1H and 13C nuclear magnetic resonance spectra were recorded on Bruker 400 MHz spectrometer in δ scale (ppm), using DMSO as solvent and TMS as the internal standard signal at Center for Drug Discovery Research and Development, Faculty of Pharmacy, Ain Shams University, Cairo, Egypt. Infra-red (IR) spectra were recorded using Shimadzu FT-IR 8400s spectrophotometer (Kyoto, Japan) and KBr discs at the Center for Drug Discovery Research and Development, Faculty of Pharmacy, Ain Shams University, Cairo, Egypt. Elemental analyses were conducted on a Thermo Scientific Flash 2000 elemental analyser (Waltham, MA) at the Regional Center for Mycology and Biotechnology, Al-Azhar University, Cairo, Egypt. High-resolution mass spectrometry analyses were performed at the Faculty of Pharmacy, Al-Fayoum University, Faiyum, Egypt, using an Agilent Q-TOF LC/MS system (Santa Clara, CA) equipped with an electrospray ionisation (ESI) source and time-of-flight (TOF) detector.
General procedure for the synthesis of N-aryl-2,5-dimethylpyrrole derivatives (IIIa–c): A mixture of appropriate aniline (2 mmol) (IIa–c) and 2,5-hexanedione (2 mmol, 0.22 g, 0.22 mL) was stirred for 1–6 h at room temperature in the presence of sulphamic acid (0.1 mmol, 0.1 g) as a catalyst. TLC monitored the reaction, and the solvent system used for TLC monitoring was ethyl acetate:hexane (3:7). Then, the reaction mixture was poured on 25 mL of ice water, and the formed precipitate was filtered and crystallised from ethanol, affording pure substituted pyrroles IIIa–c, which were used for the next step without further purification.
2,5-Dimethyl-1-phenyl-1H-pyrrole (IIIa). Buff crystals (0.33 g, 98%); MP: 51–52 °C (as reported)40,41.
1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrole (IIIb). White crystals (0.4 g, 80%); MP: 47–49 °C (as reported)43.
1-([1,1′-Biphenyl]-4-yl)-2,5-dimethyl-1H-pyrrole (IIIc). Dark red crystals (0.45 g, 92.6%); MP: 67–70 °C (as reported).42
General procedure for the synthesis of N-aryl-2,5-dimethyl-1H-pyrrole-3-carbaldehydes (IVa–c): Vilsmeier–Haack reagent was prepared by adding phosphorus oxychloride (POCl3) (3 mmol, 0.46 g, 0.28 mL) dropwise to ice-cooled DMF (10 mL). A solution of the appropriate pyrrole derivative (IIIa–c) (1 mmol) in DMF (5 mL) was added slowly to the Vilsmeier–Haack reagent. The reaction mixture was allowed to warm to room temperature, then heated to 60 °C for 4 h. The reaction progress was monitored by TLC, which showed a new spot and the disappearance of the starting material. After that, the reaction mixture was cooled to room temperature, poured into crushed ice, and quenched with 20% NaOH, forming the precipitate. The resultant precipitate was filtered and washed with water to get 2,5-dimethyl-1-aryl-1H-pyrrole-3-carbaldehydes IVa–c.
2,5-Dimethyl-1-phenyl-1H-pyrrole-3-carbaldehyde (IVa). Grey solid (0.1 g, 86%); MP: 90 °C (as reported)79. IR (KBr, cm−1): 1730 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.93 (s, 3H, CH3), 2.23 (s, 3H, CH3), 6.3 (s, 1H, pyrrolyl), 7.35 (d, J = 8 Hz, 2H, ArH), 7.52–7.6 (m, 3H, ArH), 9.8 (s, 1H, CH═O), The 1H NMR is in accordance with literature45.
1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrole-3-carbaldehyde (IVb). Brown solid (0.12 g, 60%); MP: 120 °C. IR (KBr, cm−1): 1730 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.93 (s, 3H, CH3), 2.23 (s, 3H, CH3), 6.3 (s, 1H, pyrrolyl), 7.42 (d, J = 8 Hz, 2H, ArH), 7.64 (d, J = 8 Hz, 2H, ArH), 9.8 (s, 1H, CH═O). The 1H NMR is in accordance with literature44.
1-([1,1′-Biphenyl]-4-yl)-2,5-dimethyl-1H-pyrrole-3-carbaldehyde (IVc). Grey solid (0.22 g, 80%); MP: 196–200 °C. IR (KBr, cm−1): 1730 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.99 (s, 3H, CH3), 2.29 (s, 3H, CH3), 6.3 (s, 1H, pyrrolyl), 7.42–7.51 (m, 5H, ArH), 7.76 (d, J = 8 Hz, 2H, ArH), 7.86 (d, J = 8 Hz, 2H, ArH), 9.83 (s, 1H, CH═O). Anal. Calcd for C19H17NO (275.34): C, 82.88; H, 6.22; N, 5.09; found: C, 82.63; H, 6.41; N, 5.19%.
General procedure for the synthesis of N-aryl-2,5-dimethyl-1H-pyrrol-3-yl-ethanones (IVd–e): Phosphorus oxychloride (3 mmol, 0.46 g, 0.28 mL) was added dropwise to DMA (10 mL) at 0 °C. The mixture was stirred for 30 min at room temperature. Then, one mmol of IIIa–c in DMA (20 mL) was added to the mixture at 0 °C. Subsequently, the reaction mixture was heated at reflux for 6 h, and then after cooling, ice water was added, and the mixture was neutralised with 20% NaOH. The precipitate was filtered, washed with water, and dried, affording the desired compounds IVd–e.
1-(2,5-Dimethyl-1-phenyl-1H-pyrrol-3-yl)ethanone (IVd). Greyish solid (0.16 g, 80%); MP: 80–81 °C (as reported)37. IR (KBr, cm−1): 1727 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.93 (s, 3H, CH3), 2.20 (s, 3H, CH3–C═O), 2.23 (s, 3H, CH3), 6.39 (s, 1H, pyrrolyl), 7.3 (d, J = 8 Hz, 2H, ArH), 7.52–7.59 (m, 3H, ArH). The 1H NMR is in accordance with literature39,46.
1-[1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrol-3-yl]ethanone (IVe). Yellow solid (0.14 g, 60%); MP: 81 °C, (as reported)38,39.
General procedure for the synthesis of hydrazinecarboximidamide derivatives (Va–e): To a solution of the intermediates IVa–e (1 mmol) in absolute ethanol (25 mL), 1 mL of concentrated HCl was added, followed by dry aminoguanidine HCO3 (4 mmol, 0.54). This mixture was heated under reflux for 20–24 h, and the progress of the reaction was monitored by TLC (8:2, DCM:methanol). The reaction mixture was rinsed with sodium bicarbonate solution, and the solid formed was filtered, giving the pure compounds Va–e.
2-((2,5-Dimethyl-1-phenyl-1H-pyrrol-3yl)methylene)hydrazinecarboximidamide (Va). Yellow solid (0.22 g, 90%); MP = 195–197 °C. IR (KBr, cm−1): 3473, 3403, 3303 (NH/NH2), 1656, 1582 (C═N). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.95 (s, 3H, CH3), 2.04 (s, 3H, CH3), 5.51 (s, 2H, NH2), 5.8 (s, 2H, NH), 6.24 (s, 1H, pyrrolyl), 7.27 (d, J = 8 Hz, 2H, ArH), 7.46 (t, J = 8 Hz, 1H, ArH), 7.54 (t, J = 8 Hz, 2H, ArH), 7.99 (s, 1H, CH═N). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 13.1, 13.9 (CH3s), 107, 120.9, 126.1, 126.9, 128.3, 128.6, 129.7, 138.4 (Ar-Cs), 148.6 (C═NH), 158.4 (C═N). HRMS (ESI) m/z: calcd for C14H17N5 [M + H]+ 256.1562, found 256.1562. Anal. Calcd for C14H17N5 (255.15): C, 65.86; H, 6.71; N, 27.43; found: C, 66.04; H, 6.83; N, 27.25%.
2-((1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrol-3yl)methylene)hydrazinecarboximidamide (Vb). Yellow solid (0.28 g, 98%); MP = 218–220 °C. IR (KBr, cm−1): 3445, 3335 (NH/NH2), 1670, 1621 (C═N). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.96 (s, 3H, CH3), 2.05 (s, 3H, CH3), 5.23 (s, 2H, NH2), 5.64 (s, 2H, NH), 6.24 (s, 1H, pyrrolyl), 7.33 (d, J = 8 Hz, 2H, ArH), 7.5 (d, J = 8 Hz, 2H, ArH), 7.98 (s, 1H, CH═N). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 11.2, 13 (CH3s), 105.5, 118, 128.7, 128.9, 129.8, 130.3, 133, 137 (Ar-Cs), 140.8 (C═NH), 158.7 (C═N). HRMS (ESI) m/z: calcd for C14H16ClN5 [M + H]+ 290.1176, found 290.1177. Anal. Calcd for C14H16ClN5 (289.11): C, 58.03; H, 5.57; N, 24.17; found: C, 58.2; H, 5.73; N, 22.04%.
2-((1-([1,1′-Biphenyl]-4-yl)-2,5-dimethyl-1H-pyrrol-3-yl)methylene)hydrazinecarboximidamide (Vc). Dark brown solid (0.26 g, 80%); MP = 201–204 °C. IR (KBr, cm−1): 3469, 3332 (NH/NH2), 1684, 1610 (C═N). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 2 (s, 3H, CH3), 2.13 (s, 3H, CH3), 5.21 (s, 2H, NH2, exchangeable by D2O), 5.9 (s, 2H, NH, exchangeable by D2O), 6.3 (s, 1H, pyrrolyl), 7.3–7.51 (m, 5H, ArH), 7.73 (d, J = 8 Hz, 2H, ArH), 7.8 (d, J = 8 Hz, 2H, ArH), 8.3 (s, 1H, CH═N). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 11.4, 13.1 (CH3s), 105.4, 114.7, 125.8, 127.2, 127.6, 127.9, 128.9, 129.1, 129.5, 137.3, 139.5, 140 (Ar-Cs), 141.1 (C═NH), 158.5 (C═N). HRMS (ESI) m/z: calcd for C20H21N5 [M + H]+ 332.1878, found 332.1882. Anal. Calcd for C20H21N5 (331.18): C, 72.48; H, 6.39; N, 21.13; found: C, 72.52; H, 6.48; N, 21.40%.
2-(1-(2,5-Dimethyl-1-phenyl-1H-pyrrol-3-yl)ethylidene)hydrazinecarboximidamide (Vd). Yellow solid (0.13 g, 50%); MP = 178–180 °C. IR (KBr, cm−1): 3474, 3403, 3303 (NH/NH2), 1652, 1606 (C═N). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.94 (s, 3H, CH3), 2.15 (s, 3H, CH3), 2.16 (s, 3H, CH3–C═N), 5.21 (s, 2H, NH2), 5.50 (s, 2H, NH), 6.08 (s, 1H, pyrrolyl), 7.25 (d, J = 8 Hz, 2H, ArH), 7.458 (t, J = 8 Hz, 1H, ArH), 7.53 (t, J = 8 Hz, 2H, ArH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 13.1, 13.9, 16.4 (CH3s), 107, 120.9, 126.1, 126.9, 128.3, 128.6, 129.7, 138.4 (Ar-Cs), 148.6 (C═NH), 158.4 (C═N). HRMS (ESI) m/z: calcd for C15H19N5 [M + H]+ 270.1717, found 270.1720. Anal. Calcd for C15H19N5 (269.16): C, 66.89; H, 7.11; N, 26.00; found: C, 67.05; H, 7.27; N, 26.21%.
2-(1-(1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrol-3-yl)ethylidene)hydrazinecarboximidamide (Ve). Brown solid (0.23 g, 77%); MP = 175–178 °C. IR (KBr, cm−1): 3466, 3321 (NH/NH2), 1616, 1595 (C═N). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.96 (s, 3H, CH3), 2.15 (s, 3H, CH3), 2.21 (s, 3H, CH3–C═N), 5.8 (s, 2H, NH2), 6.01 (s, 2H, NH), 6.11 (s, 1H, pyrrolyl), 7.29 (d, J = 8 Hz, 2H, ArH), 7.6 (d, J = 8 Hz, 2H, ArH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 13, 13.9, 16.4 (CH3s), 107.3, 121, 126.3, 127, 129.8, 130.5, 132.9, 137.3 (Ar-Cs), 148.5 (C═NH), 158.3 (C═N). HRMS (ESI) m/z: calcd for C15H18ClN5 [M + H]+ 304.1333, found 304.1333. Anal. Calcd C15H18ClN5 (303.13): C, 59.30; H, 5.97; N, 23.05; found: C, 59.47; H, 6.11; N, 23.32%.
General procedure for the synthesis of hydrazinecarbothioamide derivatives (VIa–e): A solution of the intermediates IVa–e (1 mmol) in absolute ethanol (20 mL) was added slowly dropwise to a solution of thiosemicarbazide hydrochloride (2 mmol, 0.25 g) and sodium acetate (4 mmol, 0.32 g) dissolved in 15–20 mL of distilled water. The reaction mixture was heated under reflux for 12 h, and the progress of the reaction was monitored by TLC (6:3:1, hexane:ethyl acetate:ethanol). The reaction mixture was rinsed with ice-cold water, and the solid formed was filtered and crystallised from ethanol, giving the pure compounds VIa–e.
2-((2,5-Dimethyl-1-phenyl-1H-pyrrol-3yl)methylene)hydrazinecarbothioamide (VIa). Yellow solid (0.25 g, 95%); MP = 243–245 °C. 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.95 (s, 3H, CH3), 2.23 (s, 3H, CH3), 6.10 (s, 2H, NH2), 6.28 (s, 1H, pyrrolyl), 7.27 (d, J = 8 Hz, 2H, ArH), 7.47 (t, J = 8 Hz, 1H, ArH), 7.54 (t, J = 8 Hz, 2H, ArH), 7.84 (s, 1H, CH═N), 9.72 (s, 1H, –NH). The 1H NMR is in accordance with literature80. Anal. Calcd C14H16N4S (272.11): C, 61.74; H, 5.92; N, 20.57; found: C, 62.02; H, 6.04; N, 20.80%.
2-((1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrol-3-yl)methylene)hydrazinecarbothioamide (VIb)81. Yellow solid (0.28 g, 96%); MP = 255 °C. 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.95 (s, 3H, CH3), 2.2 (s, 3H, CH3), 6.17 (s, 2H, NH2), 6.29 (s, 1H, pyrrolyl), 7.33 (d, J = 8 Hz, 2H, ArH), 7.33 (d, J = 8 Hz, 2H, ArH), 7.57 (d, J = 8 Hz, 2H, ArH),7.82 (s, 1H, CH═N), 9.72 (s, 1H, –NH). Anal. Calcd C14H15ClN4S (306.81): C, 54.81; H, 4.93; N, 18.26; found: C, 55.06; H, 5.12; N, 18.45%.
2-((1-([1,1′-Biphenyl]-4-yl)-2,5-dimethyl-1H-pyrrol-3-yl)methylene)hydrazinecarbothioamide (VIc). Brown solid (0.23 g, 70%); MP = 224–226 °C. IR (KBr, cm−1): 3477, 3403 (NH/NH2), 1681 (C═N), 1482 (C═S). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 2 (s, 3H, CH3), 2.2 (s, 3H, CH3), 6.17 (s, 2H, NH2), 6.34 (s, 1H, pyrrolyl), 7.38–7.53 (m, 5H, ArH), 7.73 (d, J = 8 Hz, 2H, ArH), 7.80 (d, J = 8 Hz, 2H, ArH), 8.57 (s, 1H, CH═N), 9.73 (s, 1H, –NH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 11.2, 13.1 (CH3s), 104.86, 116.3, 127.2, 128, 128.3, 128.9, 129.5, 130, 133.3, 136.7, 139.5, 140.46 (Ar-Cs), 157.4 (C═NH), 199.2 (C═S). Anal. Calcd C20H20N4S (348.14): C, 68.93; H, 5.79; N, 16.08; found: C, 69.15; H, 5.86; N, 16.27%
2-(1-(2,5-Dimethyl-1-phenyl-1H-pyrrol-3-yl)ethylidene)hydrazinecarbothioamide (VId). Buff solid (0.4 g, 84%); MP = 230–233 °C. IR (KBr, cm−1): 3466, 3342 (NH/NH2), 1684 (C═N), 1521 (C═S). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.94 (s, 3H, CH3), 2.13 (s, 3H, CH3), 2.3 (s, 3H, CH3–C═N), 6.14 (s, 2H, NH2), 6.38 (s, 1H, pyrrolyl), 7.24 (d, J = 8 Hz, 2H, ArH), 7.47 (t, J = 8 Hz, 1H, ArH), 7.54 (t, J = 8 Hz, 2H, ArH), 9.05 (s, 1H, NH).13C NMR (100 MHz, DMSO-d6) δ (ppm): 13, 13.6, 16.1 (CH3s), 107, 119.2, 127.2, 128.6, 129.8, 130, 138.1, 144.9 (Ar-Cs), 158.1 (C═NH), 194.2 (C═N). HRMS (ESI) m/z: calcd for C15H18N4S [M + H]+ 287.1334, found 287.1499. Anal. Calcd C15H18N4S (286.13): C, 62.91; H, 6.33; N, 19.56; found: C, 63.15; H, 6.45; N, 19.75%.
2-(1-(1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrol-3-yl)ethylidene)hydrazinecarbothioamide (VIe). Yellow solid (0.3 g, 78%); MP = 200–203 °C. IR (KBr, cm−1): 3463, 3382 (NH/NH2), 1695 (C═N), 1521 (C═S). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.95 (s, 3H, CH3), 2.14 (s, 3H, CH3), 2.30 (s, 3H, CH3–C═N), 6.15 (s, 2H, NH2), 6.40 (s, 1H, pyrrolyl), 7.32 (d, J = 8 Hz, 2H, ArH), 7.58 (d, J = 8 Hz, 2H, ArH), 8.98 (s, 1H, NH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 12.9, 13.5, 16.1 (CH3s), 108.8, 119.4, 127.3, 129.8, 130.4, 133.1, 137, 144.8 (Ar-Cs), 158 (C═NH), 194.3 (C═S). Anal. Calcd C15H18ClN4S (321.86): C, 55.98; H, 5.63; N, 17.41; found: C, 56.09; H, 5.41; N, 17.63%.
General procedure for the synthesis of hydrazinecarboxamide derivatives (VIIa–e): A solution of the key intermediates IVa–e (1 mmol) in absolute ethanol (20 mL) was added slowly dropwise to a solution of semicarbazide hydrochloride (0.2 mmol, 0.25 g) and sodium acetate (0.2 mmol, 0.32 g) dissolved in 15–20 mL of distilled water. The reaction mixture was heated under reflux for 12 h, and the progress of the reaction was monitored by TLC (6:3:1, hexane:ethyl acetate:ethanol). The reaction mixture was rinsed with ice-cold water, and the solid formed was filtered and crystallised from hot ethanol, giving the pure compounds VIIa–e.
2-((2,5-dimethyl-1-phenyl-1H-pyrrol-3-yl)methylene)hydrazinecarboxamide (VIIa)82. Yellow solid (0.23 g, 95%); MP = 250–255 °C (reported = 294 °C82) 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.94 (s, 3H, CH3), 2.01 (s, 3H, CH3), 6.17 (s, 2H, NH2), 6.28 (s, 1H, pyrrolyl), 7.27 (d, J = 8 Hz, 2H, ArH), 7.47 (t, J = 8 Hz, 1H, ArH), 7.54 (t, J = 8 Hz, 2H, ArH), 7.84 (s, 1H, CH═N), 9.72 (s, 1H, –NH). Anal. Calcd C14H16N4O (256.30): C, 65.61; H, 6.29; N, 21.86; found: C, 65.83; H, 6.45; N, 21.98%.
2-((1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrol-3-yl)methylene)hydrazinecarboxamide (VIIb). Yellow solid (0.26 g, 90%); MP = 248–251 °C. IR (KBr, cm−1): 3466 (NH/NH2), 1645 (C═N), 1681 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.95 (s, 3H, CH3), 2.02 (s, 3H, CH3), 6.17 (s, 2H, NH2), 6.29 (s, 1H, pyrrolyl), 7.33 (d, J = 8 Hz, 2H, ArH), 7.58 (d, J = 8 Hz, 2H, ArH), 7.82 (s, 1H, CH═N), 9.72 (s, 1H, –NH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 11.1, 13 (CH3s), 105, 116.5, 129.4, 129.9, 130.3, 133.2, 136.4, 136.7 (Ar-Cs), 146.4 (C═N), 157.3 (C═O). HRMS (ESI) m/z: calcd for C14H17ClN4O [M + H]+ 291.1015, found 291.1016. Anal. Calcd C14H17ClN4O (290.75): C, 59.11; H, 5.62; N, 18.38; found: C, 59.34; H, 3.78; N, 18.60%.
2-((1-([1,1′-Biphenyl]-4-yl)-2,5-dimethyl-1H-pyrrol-3-yl)methylene)hydrazinecarboxamide (VIIc). Brown solid (0.27 g, 84%); MP = 230–233 °C. IR (KBr, cm−1): 3466 (NH/NH2), 1613 (C═N), 1684 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 2.03 (s, 3H, CH3), 2.2 (s, 3H, CH3), 6.17 (s, 2H, NH2), 6.32 (s, 1H, pyrrolyl), 7.34–7.53 (m, 5H, ArH), 7.73 (d, J = 8 Hz, 2H, ArH), 7.8 (d, J = 8 Hz, 2H, ArH), 8.57 (s, 1H, CH═N), 9.72 (s, 1H, –NH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 11.2, 13.1(CH3s), 104.8, 116.3, 125.8, 127.2, 128, 128.3, 128.9, 129.5, 130.1, 134, 136.8, 139.5 (Ar-Cs), 140.2 (C═N), 157.4 (C═O). HRMS (ESI) m/z: calcd for C20H20N4O [M + H]+ 333.1717, found 333.1718. Anal. Calcd C20H20N4O (332.16): C, 72.04; H, 6.06; N, 16.86; found: C, 72.04; H, 6.31; N, 16.9%.
2-(1-(2,5-Dimethyl-1-phenyl-1H-pyrrol-3-yl)ethylidene)hydrazine carboxamide (VIId). Buff solid (0.21 g, 78%); MP = 238–240 °C. IR (KBr, cm−1): 3463 (NH/NH2), 1585 (C═N), 1687 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.93 (s, 3H, CH3), 2.13 (s, 3H, CH3), 2.32 (s, 3H,CH3–C═N), 6.14 (s, 2H, NH2), 6.39 (s, 1H, pyrrolyl), 7.24 (d, J = 8 Hz, 2H, ArH), 7.44 (t, J = 8 Hz, 1H, ArH), 7.51 (t, J = 8 Hz, 2H, ArH), 9.04(s, 1H, NH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 12.9, 13.6, 16.1 (CH3s), 107, 119.2, 127.1, 127.4, 128.6, 129.8, 130, 138.1 (Ar-Cs), 144.9 (C═NH), 158.1 (C═O). HRMS (ESI) m/z: calcd for C15H18N4O [M + H]+ 271.1559, found 271.1560. Anal. Calcd C15H18N4O (270.15): C, 66.64; H, 6.71; N, 20.73; found: C, 66.81; H, 6.85; N, 20.97%.
2-(1-(1-(4-Chlorophenyl)-2,5-dimethyl-1H-pyrrol-3-yl)ethylidene)hydrazinecarboxamide (VIIe). Yellow solid (0.19 g, 65%); MP = 244–247 °C. IR (KBr, cm−1): 3466 (NH/NH2), 1578 (C═N), 1687 (C═O). 1H NMR (400 MHz, DMSO-d6) δ (ppm): 1.95 (s, 3H, CH3), 2.14 (s, 3H, CH3), 2.5 (s, 3H, CH3–C═N), 6.15 (s, 2H, NH2), 6.40 (s, 1H, pyrrolyl), 7.32 (d, J = 8 Hz, 2H, ArH), 7.58 (d, J = 8 Hz, 2H, ArH), 8.98 (s, 1H, NH). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 12.8, 13.5, 16.1 (CH3s), 108.7, 120.7, 127.5, 128.5, 130, 130.4, 133.8, 134.7 (Ar-Cs), 136 (C═NH), 158 (C═S). HRMS (ESI) m/z: calcd for C15H17ClN4O [M + H]+ 305.1171, found 305.1173. Anal. Calcd C15H17ClN4O (304.77): C, 59.11; H, 5.62; N, 18.38; found: C, 59.35; H, 5.75; N, 18.62%.
Biological evaluation
Antimicrobial activity
The minimum inhibitory concentrations (MICs) of the synthesised final compounds and control antibiotics were determined using the broth microdilution method according to the guidelines outlined by the Clinical and Laboratory Standards Institute (CLSI)83 or as described in previous reports84,85.
In vitro cytotoxic assay
Compounds (Vb, Vc, Vd, and Ve) were assayed against normal mammalian cell line VERO (African Green Monkey Kidney cells) to determine the potential toxic effect on mammalian cells in vitro, as described in previous reports86,87.
Antibiofilm assay
The biofilm formation was quantified as previously described88–90. Briefly, bacterial overnight cultures in tryptone soya broth (TSB) were prepared and diluted with TSB to an optical density OD600 nm of 0.2 in 96 wells microtiter plate (150 µL per well). 1/4th MIC dose (dissolved in DMSO) of each drug was added to each well of the tested bacterial species. One well representing negative control (bacteria free) and another well representing positive control (with vehicle) were included. The plates were incubated for 24 h at 37 °C. The planktonic cells were aspirated, and the plates were washed softly with deionised water and then left to dry. The adherent bacterial cells were fixed for 25 min with methanol, followed by staining with crystal violet (1% w/v) for another 20 min. The excessive dye was washed, dried, then stained biofilms were imaged using a fully motorised Leica AF7000 microscope with a 20× objective lens. Finally, the adhered dye was dissolved in glacial acetic acid (33% v/v), and the absorbance was measured at 590 nm using a Biotek Spectrofluorometer (Biotek, Winooski, VT). The test was made in triplicate and the absorbance was shown as mean ± standard error of percentage changes from untreated controls. The percentage inhibition and reduction of biofilm growth was calculated by using the following equation:
In the above equation, Nc stands for negative control, and E for experimental samples.
Antiprotease activity
To assess the inhibitory effect of the tested drugs on bacterial protease activity, the skim milk agar method was employed according to previous descriptions88–90. Suspensions from overnight cultures in 96-well plate of the indicated bacteria in TSB, along with each drug (1/4th of the MIC) were centrifuged at 10 000 × g for 20 min to pellet the planktonic cells. Subsequently, aliquots (100 µL) of supernatants were added to wells made in 5% skim milk agar plates, followed by incubation at 37 °C overnight. Supernatants from untreated cultures, and 1% SDS were used as positive control for protease activity (clear zone around the well). Negative control of protease activity was bacteria free media.
In silico studies
Docking studies
All the molecular docking studies done were based on the crystal structure of undecaprenyl diphosphate phosphatase (UPPP) (PDB code: 6CB2)64. Molecular docking was done via the CDOCKER protocol DS. The CDOCKER protocol allows us to simulate the docking of a ligand into the target’s binding site and utilises several scoring functions to assess the docked poses91. It has been demonstrated that the CHARMM-based CDOCKER protocol yields highly accurate docked poses92. The protein preparation tool was used to correct common problems in the protein structure by adding missing loops and hydrogens and excluding alternate conformers. All water molecules that did not contribute to the binding interactions were removed. The binding site was then identified using the define and edit binding site tool, resulting in a sphere of 14 Å, and the binding site atomic co-ordinates were X: 0.499, Y: 4.206, and Z: 6.943.
The reference lead compound (E)-2-(1-(2-(4-butylphenyl)-4-methylthiazol-5-yl)ethylidene)hydrazinecarboximidamide and our N-arylpyrrole designed derivatives were sketched using DS and then prepared by the ligand preparation tool to fix any incorrect valences and generate 3D conformers. Both the reference ligand and our compounds were docked into the UPPP binding site, and the reference ligand interactions with UPPP were used to validate and justify the interactions of our newly designed compounds.
ADMET studies
In silico studies for the prediction of physicochemical/pharmacokinetic features and toxicity of the most potent compounds (Vb, Vc, Vd, and Ve) have been performed using the ADMET descriptors algorithm in Discovery Studio 2024 software package29,67 to investigate their ADMET profiles.
Supplementary Material
Acknowledgements
The authors express thanks to the Center of Excellence of Drug Discovery and Development Research at Misr International University. We acknowledge support of the publication fee by the CSIC Open Access Publication Support Initiative through its Unit of Information Resources for Research (URICI).
Funding Statement
The researchers did not obtain any funding for the work they submitted. The publishing fee is covered under the CSIC agreement with Taylor & Francis. GY was supported by AYUDAS JUAN DE LA CIERVA-INCORPORACIÓN (IJC2020-044166-I). This work is also supported by the Ministerio de Ciencia, Innovación y Universidades Spain.
Disclosure statement
The authors report no conflicts of interest.
References
- 1.Rossolini GM, Arena F, Pecile P, Pollini S.. Update on the antibiotic resistance crisis. Curr Opin Pharmacol. 2014;18:56–60. [DOI] [PubMed] [Google Scholar]
- 2.de Kraker ME, Stewardson AJ, Harbarth S.. Will 10 million people die a year due to antimicrobial resistance by 2050? PLoS Med. 2016;13(11):e1002184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Mulani MS, Kamble EE, Kumkar SN, Tawre MS, Pardesi KR.. Emerging strategies to combat ESKAPE pathogens in the era of antimicrobial resistance: a review. Front Microbiol. 2019;10:539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.El Malah T, Nour HF, Satti AAE, Hemdan BA, El-Sayed WA.. Design, synthesis, and antimicrobial activities of 1, 2, 3-triazole glycoside clickamers. Molecules. 2020;25(4):790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Voss A, Doebbeling BN.. The worldwide prevalence of methicillin-resistant Staphylococcus aureus. Int J Antimicrob Agents. 1995;5(2):101–106. [DOI] [PubMed] [Google Scholar]
- 6.Stapleton PD, Taylor PW.. Methicillin resistance in Staphylococcus aureus: mechanisms and modulation. Sci Prog. 2002;85(Pt 1):57–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Costerton JW, Stewart PS, Greenberg EP.. Bacterial biofilms: a common cause of persistent infections. Science. 1999;284(5418):1318–1322. [DOI] [PubMed] [Google Scholar]
- 8.Cabot G, Zamorano L, Moyà B, Juan C, Navas A, Blázquez J, Oliver A.. Evolution of Pseudomonas aeruginosa antimicrobial resistance and fitness under low and high mutation rates. Antimicrob Agents Chemother. 2016;60(3):1767–1778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Oman HS, Kharat AA, Phatak PS, Haval KP, Kulkarni JA, Kakde GS, Kharat KR, Kadam DG, Kharat AS.. Synthesis and evaluation of 1,2,3-triazole benzoate derivatives for inhibition of serine β-lactamases in extensively drug resistant pathogenic E. coli strains. Eur J Med Chem Rep. 2023;9:100109. [Google Scholar]
- 10.Lee C-R, Lee JH, Park M, Park KS, Bae IK, Kim YB, Cha C-J, Jeong BC, Lee SH.. Biology of Acinetobacter baumannii: pathogenesis, antibiotic resistance mechanisms, and prospective treatment options. Front Cell Infect Microbiol. 2017;7:55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Munoz-Price LS, Poirel L, Bonomo RA, Schwaber MJ, Daikos GL, Cormican M, Cornaglia G, Garau J, Gniadkowski M, Hayden MK, et al. Clinical epidemiology of the global expansion of Klebsiella pneumoniae carbapenemases. Lancet Infect Dis. 2013;13(9):785–796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Redgrave LS, Sutton SB, Webber MA, Piddock LJV.. Fluoroquinolone resistance: mechanisms, impact on bacteria, and role in evolutionary success. Trends Microbiol. 2014;22(8):438–445. [DOI] [PubMed] [Google Scholar]
- 13.Park SH. Third-generation cephalosporin resistance in gram-negative bacteria in the community: a growing public health concern. Korean J Intern Med. 2014;29(1):27–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Idhayadhulla A, Surendra Kumar R, Abdul Nasser AJ, Manilal A.. Synthesis and antimicrobial activity of some new pyrrole derivatives. Bull Chem Soc Eth. 2012;26(3):429–435. [Google Scholar]
- 15.Masci D, Hind C, Islam MK, Toscani A, Clifford M, Coluccia A, Conforti I, Touitou M, Memdouh S, Wei X, et al. Switching on the activity of 1,5-diaryl-pyrrole derivatives against drug-resistant ESKAPE bacteria: structure–activity relationships and mode of action studies. Eur J Med Chem. 2019;178:500–514. [DOI] [PubMed] [Google Scholar]
- 16.Biava M, Porretta GC, Poce G, Supino S, Deidda D, Pompei R, Molicotti P, Manetti F, Botta M.. Antimycobacterial agents. Novel diarylpyrrole derivatives of BM212 endowed with high activity toward Mycobacterium tuberculosis and low cytotoxicity. J Med Chem. 2006;49(16):4946–4952. [DOI] [PubMed] [Google Scholar]
- 17.Mohammad H, Mayhoub AS, Ghafoor A, Soofi M, Alajlouni RA, Cushman M, Seleem MN.. Discovery and characterization of potent thiazoles versus methicillin-and vancomycin-resistant Staphylococcus aureus. J Med Chem. 2014;57(4):1609–1615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Hagras M, Mohammad H, Mandour MS, Hegazy YA, Ghiaty A, Seleem MN, Mayhoub AS.. Investigating the antibacterial activity of biphenylthiazoles against methicillin-and vancomycin-resistant Staphylococcus aureus (MRSA and VRSA). J Med Chem. 2017;60(9):4074–4085. [DOI] [PubMed] [Google Scholar]
- 19.Mohammad H, Younis W, Chen L, Peters CE, Pogliano J, Pogliano K, Cooper B, Zhang J, Mayhoub A, Oldfield E, et al. Phenylthiazole antibacterial agents targeting cell wall synthesis exhibit potent activity in vitro and in vivo against vancomycin-resistant Enterococci. J Med Chem. 2017;60(6):2425–2438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Asghar SF, Yasin KA, Aziz S.. Synthesis and cyclisation of 1,4-disubstituted semicarbazides. Nat Prod Res. 2010;24(4):315–325. [DOI] [PubMed] [Google Scholar]
- 21.Rane RA, Naphade SS, Bangalore PK, Palkar MB, Shaikh MS, Karpoormath R.. Synthesis of novel 4-nitropyrrole-based semicarbazide and thiosemicarbazide hybrids with antimicrobial and anti-tubercular activity. Bioorg Med Chem Lett. 2014;24(14):3079–3083. [DOI] [PubMed] [Google Scholar]
- 22.Salgın-Gökşen U, Gökhan-Kelekçi N, Göktaş Ö, Köysal Y, Kılıç E, Işık Ş, Aktay G, Özalp M.. 1-Acylthiosemicarbazides, 1,2,4-triazole-5(4H)-thiones, 1,3,4-thiadiazoles and hydrazones containing 5-methyl-2-benzoxazolinones: synthesis, analgesic-anti-inflammatory and antimicrobial activities. Bioorg Med Chem. 2007;15(17):5738–5751. [DOI] [PubMed] [Google Scholar]
- 23.Kasare SL, Gund PN, Sathe BP, Patil PS, Rehman NNMA, Dixit PP, Choudhari PB, Haval KP.. Synthesis, antimicrobial screening, and docking study of new 2‐(2‐ethylpyridin‐4‐yl)‐4‐methyl‐N‐phenylthiazole‐5‐carboxamide derivatives. J Chin Chem Soc. 2021;68(2):353–361. [Google Scholar]
- 24.Sakkiah S, Lee KW.. Pharmacophore-based virtual screening and density functional theory approach to identifying novel butyrylcholinesterase inhibitors. Acta Pharmacol Sin. 2012;33(7):964–978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Debnath AK. Pharmacophore mapping of a series of 2,4-diamino-5-deazapteridine inhibitors of Mycobacterium avium complex dihydrofolate reductase. J Med Chem. 2002;45(1):41–53. [DOI] [PubMed] [Google Scholar]
- 26.Hammad A, Abutaleb NS, Elsebaei MM, Norvil AB, Alswah M, Ali AO, Abdel-Aleem JA, Alattar A, Bayoumi SA, Gowher H, et al. From phenylthiazoles to phenylpyrazoles: broadening the antibacterial spectrum toward carbapenem-resistant bacteria. J Med Chem. 2019;62(17):7998–8010. [DOI] [PubMed] [Google Scholar]
- 27.Shahin IG, Abutaleb NS, Alhashimi M, Kassab AE, Mohamed KO, Taher AT, Seleem MN, Mayhoub AS.. Evaluation of N-phenyl-2-aminothiazoles for treatment of multi-drug resistant and intracellular Staphylococcus aureus infections. Eur J Med Chem. 2020;202:112497. [DOI] [PubMed] [Google Scholar]
- 28.Omara M, Hagras M, Elsebaie MM, Abutaleb NS, Nour El-Din HT, Mekhail MO, Attia AS, Seleem MN, Sarg MT, Mayhoub AS, et al. Exploring novel aryl/heteroaryl-isosteres of phenylthiazole against multidrug-resistant bacteria. RSC Adv. 2023;13(29):19695–19709. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Gaurav A, Gautam V.. Structure-based three-dimensional pharmacophores as an alternative to traditional methodologies. J Recep Ligand Channel Res. 2014;7:27–38. [Google Scholar]
- 30.Arooj M, Thangapandian S, John S, Hwang S, Park JK, Lee KW.. 3D QSAR pharmacophore modeling, in silico screening, and density functional theory (DFT) approaches for identification of human chymase inhibitors. Int J Mol Sci. 2011;12(12):9236–9264. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Kim H-J, Doddareddy MR, Choo H, Cho YS, No KT, Park W-K, Pae AN.. New serotonin 5-HT6 ligands from common feature pharmacophore hypotheses. J Chem Inf Model. 2008;48(1):197–206. [DOI] [PubMed] [Google Scholar]
- 32.Pal S, Kumar V, Kundu B, Bhattacharya D, Preethy N, Reddy MP, Talukdar A.. Ligand-based pharmacophore modeling, virtual screening and molecular docking studies for discovery of potential topoisomerase I inhibitors. Comput Struct Biotechnol J. 2019;17:291–310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Qandeel BM, Mowafy S, Abouzid K, Farag NA.. Lead generation of UPPS inhibitors targeting MRSA: using 3D-QSAR pharmacophore modeling, virtual screening, molecular docking, and molecular dynamic simulations. BMC Chem. 2024;18(1):14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.De SK. Sulfamic acid as a novel, efficient, cost‐effective, and reusable solid acid catalyst for the synthesis of pyrroles under solvent‐free conditions. Synth Commun. 2008;38(5):803–809. [Google Scholar]
- 35.Abdelhamid IA, Shaaban MR, Elwahy AH. Applications of the Vilsmeier reaction in heterocyclic chemistry. In: Scriven EFV, Ramsden CA, editors. Advances in heterocyclic chemistry. Amsterdam: Elsevier; 2022. p. 171–223. [Google Scholar]
- 36.Rajanna KC, Moazzam Ali M, Sana S, Saiprakash PK.. Vilsmeier Haack acetylation in micellar media: an efficient one pot synthesis of 2‐chloro‐3‐acetyl quinolines. J Dispers Sci Technol. 2004;25(1):17–21. [Google Scholar]
- 37.Arcadi A, Giuseppe SD, Marinelli F, Rossi E.. Gold‐catalyzed sequential amination/annulation reactions of 2‐propynyl‐1,3‐dicarbonyl compounds. Adv Synth Catal. 2001;343(5):443–446. [Google Scholar]
- 38.Wang Y, Bi X, Li D, Liao P, Wang Y, Yang J, Zhang Q, Liu Q.. Iron-catalyzed synthesis of polysubstituted pyrroles via [4C + 1N] cyclization of 4-acetylenic ketones with primary amines. Chem Commun. 2011;47(2):809–811. [DOI] [PubMed] [Google Scholar]
- 39.Chen X, Yuan J, Zhou M.. Efficient synthesis of substituted pyrroles through iron-mediated oxidation reaction of 3-methylenehexane-2,5-dione with primary amines. Synth Commun. 2019;49(1):32–38. [Google Scholar]
- 40.Schirmacher RH, Rösch D, Thomas F.. Hexafluoroisopropanol as solvent and promotor in the Paal–Knorr synthesis of N-substituted diaryl pyrroles. Tetrahedron. 2021;83:131985. [Google Scholar]
- 41.Mohsenzadeh F, Darabi HR, Alivand M, Aghapoor K, Balavar Y.. Naturally occurring organic acids for organocatalytic synthesis of pyrroles via Paal–Knorr reaction. Res Chem Intermed. 2020;46(12):5255–5262. [Google Scholar]
- 42.Stanley J, Hazlewood GK, Hughes, F.. Pyrroles derived from acetonylacetone. J Proc R Soc New South Wales. 1937;71:92–101. [Google Scholar]
- 43.Shokri F, Behbahani FK.. Synthesis of Fe3O4@l-proline@ SO3H as a novel and reusable acidic magnetic nanocatalyst and its application for the synthesis of N-substituted pyrroles at room temperature under ultrasonic irradiation and without solvent. Inorg Nano-Met Chem. 2022;52(8):1143–1152. [Google Scholar]
- 44.Bhakta S, Scalacci N, Maitra A, Brown AK, Dasugari S, Evangelopoulos D, McHugh TD, Mortazavi PN, Twist A, Petricci E, et al. Design and synthesis of 1-((1,5-bis(4-chlorophenyl)-2-methyl-1H-pyrrol-3-yl) methyl)-4-methylpiperazine (BM212) and N-Adamantan-2-yl-N′-((E)-3,7-dimethylocta-2,6-dienyl)ethane-1,2-diamine (SQ109) pyrrole hybrid derivatives: discovery of potent antitubercular agents effective against multidrug-resistant mycobacteria. J Med Chem. 2016;59(6):2780–2793. [DOI] [PubMed] [Google Scholar]
- 45.Oderinlo OO, Jordaan A, Seldon R, Isaacs M, Hoppe HC, Warner DF, Tukulula M, Khanye SD.. Hydrazone‐tethered 5‐(pyridin‐4‐yl)‐4H‐1,2,4‐triazole‐3‐thiol hybrids: synthesis, characterisation, in silico ADME studies, and in vitro antimycobacterial evaluation and cytotoxicity. ChemMedChem. 2023;18(6):e202200572. [DOI] [PubMed] [Google Scholar]
- 46.Kashid AM, Dube PN, Alkutkar PG, Bothara KG, Mokale SN, Dhawale SC.. Synthesis, biological screening and ADME prediction of benzylindole derivatives as novel anti-HIV-1, anti-fungal and anti-bacterial agents. Med Chem Res. 2013;22(10):4633–4640. [Google Scholar]
- 47.Ciaccia M, Di Stefano S.. Mechanisms of imine exchange reactions in organic solvents. Org Biomol Chem. 2015;13(3):646–654. [DOI] [PubMed] [Google Scholar]
- 48.El-Telbany M, Mohamed AA, Yahya G, Abdelghafar A, Abdel-Halim MS, Saber S, Alfaleh MA, Mohamed AH, Abdelrahman F, Fathey HA, et al. Combination of meropenem and zinc oxide nanoparticles; antimicrobial synergism, exaggerated antibiofilm activity, and efficient therapeutic strategy against bacterial keratitis. Antibiotics. 2022;11(10):1374. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Singh S, Datta S, Narayanan KB, Rajnish KN.. Bacterial exo-polysaccharides in biofilms: role in antimicrobial resistance and treatments. J Genet Eng Biotechnol. 2021;19(1):140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Culp E, Wright GD.. Bacterial proteases, untapped antimicrobial drug targets. J Antibiot. 2017;70(4):366–377. [DOI] [PubMed] [Google Scholar]
- 51.Helmerhorst EJ, Reijnders IM, van ‘t Hof W, Veerman EC, Nieuw Amerongen AV.. A critical comparison of the hemolytic and fungicidal activities of cationic antimicrobial peptides. FEBS Lett. 1999;449(2–3):105–110. [DOI] [PubMed] [Google Scholar]
- 52.Palchykov V, Manko N, Finiuk N, Stoika R, Obushak M, Pokhodylo N. Antimicrobial action of arylsulfonamides bearing (aza)norbornane and related motifs: evaluation of new promising anti-MRSA agents. Med Chem Res. 2022:1–9. [Google Scholar]
- 53.Famuyide IM, Aro AO, Fasina FO, Eloff JN, McGaw LJ.. Antibacterial and antibiofilm activity of acetone leaf extracts of nine under-investigated South African Eugenia and Syzygium (Myrtaceae) species and their selectivity indices. BMC Complement Altern Med. 2019;19(1):141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.López-Lázaro M. How many times should we screen a chemical library to discover an anticancer drug? Drug Discov Today. 2015;20(2):167–169. [DOI] [PubMed] [Google Scholar]
- 55.Indrayanto G, Putra GS, Suhud F.. Validation of in-vitro bioassay methods: application in herbal drug research. Profiles Drug Subst Excip Relat Methodol. 2021;46:273–307. [DOI] [PubMed] [Google Scholar]
- 56.Peña-Morán OA, Villarreal ML, Álvarez-Berber L, Meneses-Acosta A, Rodríguez-López V.. Cytotoxicity, post-treatment recovery, and selectivity analysis of naturally occurring podophyllotoxins from Bursera fagaroides var. fagaroides on breast cancer cell lines. Molecules. 2016;21(8):1013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Nogueira F, do Rosário V.. Methods for assessment of antimalarial activity in the different phases of the Plasmodium life cycle. Rev Pan-Amaz Saude. 2010;1(3):109–124. [Google Scholar]
- 58.Eid I, Elsebaei MM, Mohammad H, Hagras M, Peters CE, Hegazy YA, Cooper B, Pogliano J, Pogliano K, Abulkhair HS, et al. Arylthiazole antibiotics targeting intracellular methicillin-resistant Staphylococcus aureus (MRSA) that interfere with bacterial cell wall synthesis. Eur J Med Chem. 2017;139:665–673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Goldman R, Strominger JL.. Purification and properties of C55-isoprenylpyrophosphate phosphatase from Micrococcus lysodeikticus. J Biol Chem. 1972;247(16):5116–5122. [PubMed] [Google Scholar]
- 60.Czarny TL, Brown ED.. A small-molecule screening platform for the discovery of inhibitors of undecaprenyl diphosphate synthase. ACS Infect Dis. 2016;2(7):489–499. [DOI] [PubMed] [Google Scholar]
- 61.Inokoshi J, Nakamura Y, Komada S, Komatsu K, Umeyama H, Tomoda H.. Inhibition of bacterial undecaprenyl pyrophosphate synthase by small fungal molecules. J Antibiot. 2016;69(11):798–805. [DOI] [PubMed] [Google Scholar]
- 62.Koyama N, Inokoshi J, Tomoda H.. Anti-infectious agents against MRSA. Molecules. 2012;18(1):204–224. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Desai J, Wang Y, Wang K, Malwal SR, Oldfield E.. Isoprenoid biosynthesis inhibitors targeting bacterial cell growth. ChemMedChem. 2016;11(19):2205–2215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Workman SD, Worrall LJ, Strynadka NC.. Crystal structure of an intramembranal phosphatase central to bacterial cell-wall peptidoglycan biosynthesis and lipid recycling. Nat Commun. 2018;9(1):1159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Ezz Eldin RR, Saleh MA, Alotaibi MH, Alsuair RK, Alzahrani YA, Alshehri FA, Mohamed AF, Hafez SM, Althoqapy AA, Khirala SK, et al. Ligand-based design and synthesis of N′-benzylidene-3,4-dimethoxybenzohydrazide derivatives as potential antimicrobial agents; evaluation by in vitro, in vivo, and in silico approaches with SAR studies. J Enzyme Inhib Med Chem. 2022;37(1):1098–1119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Amin NH, El-Saadi MT, Ibrahim AA, Abdel-Rahman HM.. Design, synthesis and mechanistic study of new 1,2,4-triazole derivatives as antimicrobial agents. Bioorg Chem. 2021;111:104841. [DOI] [PubMed] [Google Scholar]
- 67.Jejurikar BL, Rohane SH.. Drug designing in discovery studio. Asian J Res Chem. 2021;14:135–138. [Google Scholar]
- 68.Cheng A, Merz KM.. Prediction of aqueous solubility of a diverse set of compounds using quantitative structure–property relationships. J Med Chem. 2003;46(17):3572–3580. [DOI] [PubMed] [Google Scholar]
- 69.Wang J, Hou T.. Recent advances on aqueous solubility prediction. Comb Chem High Throughput Screen. 2011;14(5):328–338. [DOI] [PubMed] [Google Scholar]
- 70.Kelder J, Grootenhuis PD, Bayada DM, Delbressine LP, Ploemen JP.. Polar molecular surface as a dominating determinant for oral absorption and brain penetration of drugs. Pharm Res. 1999;16(10):1514–1519. [DOI] [PubMed] [Google Scholar]
- 71.Susnow RG, Dixon SL.. Use of robust classification techniques for the prediction of human cytochrome P450 2D6 inhibition. J Chem Inf Comput Sci. 2003;43(4):1308–1315. [DOI] [PubMed] [Google Scholar]
- 72.Egan WJ, Lauri G.. Prediction of intestinal permeability. Adv Drug Deliv Rev. 2002;54(3):273–289. [DOI] [PubMed] [Google Scholar]
- 73.Colmenarejo G, Alvarez-Pedraglio A, Lavandera J-L.. Cheminformatic models to predict binding affinities to human serum albumin. J Med Chem. 2001;44(25):4370–4378. [DOI] [PubMed] [Google Scholar]
- 74.Vora J, Patel S, Sinha S, Sharma S, Srivastava A, Chhabria M, Shrivastava N.. Molecular docking, QSAR and ADMET based mining of natural compounds against prime targets of HIV. J Biomol Struct Dyn. 2019;37(1):131–146. [DOI] [PubMed] [Google Scholar]
- 75.Padariya M, Kalathiya U, Baginski M.. Docking simulations, molecular properties and ADMET studies of novel chromane-6,7-diol analogues as potential inhibitors of mushroom tyrosinase. Gene Ther Mol Biol. 2014;16:201–218. [Google Scholar]
- 76.Hosen S, Dash R, Khatun M, Akter R, Bhuiyan M, Karim M, Mouri N, Ahamed F, Islam K, Afrin S, et al. In silico ADME/T and 3D QSAR analysis of KDR inhibitors. J Appl Pharm Sci. 2017;7(1):120–128. [Google Scholar]
- 77.Zhang Q, Zheng L, Luo D, Zhao M.. In vitro simulated gastrointestinal digestion stability of a neuroprotective octapeptide WCPFSRSF and prediction of potential bioactive peptides in its digestive fragments by multiple bioinformatics tools. J Agric Food Chem. 2023;71(18):6987–6998. [DOI] [PubMed] [Google Scholar]
- 78.Khan T, Lawrence AJ, Azad I, Raza S, Joshi S, Khan AR.. Computational drug designing and prediction of important parameters using in silico methods—a review. Curr Comput Aided Drug Des. 2019;15(5):384–397. [DOI] [PubMed] [Google Scholar]
- 79.Mital A, Murugesan D, Kaiser M, Yeates C, Gilbert IH.. Discovery and optimisation studies of antimalarial phenotypic hits. Eur J Med Chem. 2015;103:530–538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Sowmya PV, Poojary B, Kumar V, Vishwanatha U, Shetty P. Fluorinated pyrrole incorporated 2-thiazolyl hydrazone motifs: a new class of antimicrobial and anti tuberculosis agents. Arch Pharm Res. 2017:1–21. [DOI] [PubMed] [Google Scholar]
- 81.Barsanti PA. Thiosemicarbazones as anti-virals and immunopotentiators. Google Patents. 2009. [Google Scholar]
- 82.Rips R, Buu-Hoï NP.. Aldehydes derived from 1,2,5-trisubstituted pyrroles. J Org Chem. 1959;24(3):372–374. [Google Scholar]
- 83.Clinical and Laboratory Standards Institute . Methods for dilution antimicrobial susceptibility tests for bacteria that grow aerobically. Approve Standard M7-A7. 7th ed. PA, USA: CLSI; 2006. [Google Scholar]
- 84.Qaiyumi S. Macro- and microdilution methods of antimicrobial susceptibility testing. In: Schwalbe R, Steele-Moore L, Goodwin AC, editors. Antimicrobial susceptibility testing protocols. Boca Raton: Taylor & Francis; 2007. p. 75–79. [Google Scholar]
- 85.Jenkins SG, Schuetz AN.. Current concepts in laboratory testing to guide antimicrobial therapy. Mayo Clin Proc. 2012;87(3):290–308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Mosmann T. Rapid colorimetric assay for cellular growth and survival: application to proliferation and cytotoxicity assays. J Immunol Methods. 1983;65(1–2):55–63. [DOI] [PubMed] [Google Scholar]
- 87.Riyadh SM, Gomha S, Mahmmoud EA, Elaasser MM.. Synthesis and anticancer activities of thiazoles, 1,3-thiazines, and thiazolidine using chitosan-grafted-poly(vinylpyridine) as basic catalyst. Heterocycles. 2015;91(6):1227–1243. [Google Scholar]
- 88.Fekry M, Yahya G, Osman A, Al-Rabia MW, Mostafa I, Abbas HA.. GC–MS analysis and microbiological evaluation of caraway essential oil as a virulence attenuating agent against Pseudomonas aeruginosa. Molecules. 2022;27(23):8532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Khan AU, Hussain T, Khan MA, Almostafa MM, Younis NS, Yahya G.. Antibacterial and antibiofilm activity of Ficus carica-mediated calcium oxide (CaONPs) phyto-nanoparticles. Molecules. 2023;28(14):5553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Rahman AU, Abdullah A, Faisal S, Mansour B, Yahya G.. Unlocking the therapeutic potential of Nigella sativa extract: phytochemical analysis and revealing antimicrobial and antioxidant marvels. BMC Complement Med Ther. 2024;24(1):266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Mhetre UV, Haval NB, Bondle GM, Rathod SS, Choudhari PB, Kumari J, Sriram D, Haval KP.. Design, synthesis and molecular docking study of novel triazole–quinazolinone hybrids as antimalarial and antitubercular agents. Bioorg Med Chem Lett. 2024;108:129800. [DOI] [PubMed] [Google Scholar]
- 92.Bakale RD, Phatak PS, Rathod SS, Choudhari PB, Rekha EM, Sriram D, Kulkarni RS, Haval KP.. In vitro and in silico exploration of newly synthesized triazolyl-isonicotinohydrazides as potent antitubercular agents. J Biomol Struct Dyn. 2025;43(3):1372–1391. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.














