Next Article in Journal
Repurposing 1,4-Dihydropyridine Scaffold: 4-Imidazo[2,1-b]thiazole-Derivatives from Calcium Entry Blockers to a New Approach for Gut Dysfunctional Motility
Previous Article in Journal
In Silico Design and Computational Elucidation of Hypothetical Resveratrol–Curcumin Hybrids as Potential Cancer Pathway Modulators
Previous Article in Special Issue
Promising Norlabdane-Heterocyclic Hybrids: Synthesis, Structural Characterization and Antimicrobial Activity Evaluation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Innovative Amino-Functionalization of Pyrido[2,3-d]pyrimidine Scaffolds for Broad Therapeutic Applications Supported by Computational Analyses

1
Basic Science Department (Chemistry), Thebes Higher Institute for Engineering, Thebes Academy, Maadi 11434, Egypt
2
Chemistry Department, Faculty of Science, Menoufia University, Shebin El-Kom 32511, Egypt
3
Pharmaceutical Medicinal Chemistry and Drug Design Department, Faculty of Pharmacy (Girls), Al-Azhar University, Cairo 11754, Egypt
4
The National Committee of Drugs, Academy of Scientific Research and Technology, Cairo 11516, Egypt
5
Microbial Genetics Department, Biotechnology Research Institute, National Research Centre, Giza 12622, Egypt
6
Egypt Center for Research and Regenerative Medicine (ECRRM), Cairo 11517, Egypt
7
Department of Pharmaceutical Sciences, College of Pharmacy, Princess Nourah Bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
8
Chemistry Department, College of Science, Jouf University, P.O. Box 2014, Sakaka, Aljouf 72341, Saudi Arabia
9
Department of Pharmaceutical Manufacturing, College of Pharmaceutical Sciences and Drug Manufacturing, Misr University for Science and Technology (MUST), 6th of October City, P.O. Box 77, Giza 12591, Egypt
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2025, 18(10), 1472; https://doi.org/10.3390/ph18101472
Submission received: 24 August 2025 / Revised: 20 September 2025 / Accepted: 22 September 2025 / Published: 30 September 2025

Abstract

Background: Derivatives of Pyrido[2,3-d]pyrimidine-6-carboxylate are promising multi-target scaffolds. This study focused on synthesizing 16 amino-functionalized derivatives and evaluating their dual anticancer and antibacterial activities, supported by mechanistic and computational analyses. Objectives: Design and synthesize derivatives, evaluate cytotoxicity against HeLa, HepG-2, and MCF-7 (selectivity against WI-38), investigate EGFRWT and EGFRT790M inhibition, assess cell cycle, apoptosis, and migration effects, antibacterial efficacy against E. coli and P. aeruginosa, and perform in silico ADMET, docking, molecular dynamics, DFT, and antiviral predictions. Methods: Synthesized 16 derivatives; tested for cytotoxicity, EGFR inhibition, cell cycle, apoptosis, migration; assessed antibacterial activity; performed ADMET profiling, molecular docking, molecular dynamics, and DFT calculations. Results: Derivatives 1, 2, and 7 showed highest cytotoxicity (IC50 = 3.98–17.52 μM; WI-38 IC50 = 64.07–81.65 μM). Compound 1 potently inhibited EGFRWT (IC50 = 0.093 μM) and EGFRT790M (IC50 = 0.174 μM), induced G0/G1 arrest (74.86%) and apoptosis (26.37%), and reduced MCF-7 migration (69.63%). Moderate antibacterial activity observed (MIC = 50 μg/mL). ADMET indicated favorable pharmacokinetics, low CYP inhibition, negative mutagenicity, and oral toxicity class III. Molecular dynamics confirmed stable binding (EGFRWT RMSD 3 Å; EGFRT790M 3.5–4.6 Å) with persistent hydrogen bonds. In silico antiviral evaluation suggested strong binding to HCV NS5A (–9.36 kcal/mol), SARS-CoV-2 Mpro (–9.82 kcal/mol), and E.coli DNA gyrase (–10.25 kcal/mol). Conclusions: Compound 1 exhibits dual anticancer and antibacterial activity, supported by mechanistic and computational analyses, highlighting pyrido[2,3-d]pyrimidines as promising multi-target therapeutic scaffolds.

1. Introduction

One of the major challenges to worldwide health is posed by cancer, which ranks second in terms of causes of death behind heart disease. Approximately 10 million people died from cancer worldwide, while 19.3 million new cases of cancer diagnoses were reported in 2023 alone, according to data from the World Health Organization and GLOBOCAN [1,2,3]. Current anticancer medications have serious drawbacks, including high toxicity, poor selectivity, drug resistance, and low absorption, despite recent advances in early diagnosis and chemotherapy [4,5,6]. These issues highlight the ongoing need for innovative therapies that focus on particular genetic factors which are involved in the genesis of cancer.
A central focus of targeted cancer therapy is EGFR, an ErbB family receptor tyrosine kinase widely recognized for its clinical relevance. Alterations in EGFR, either through gene amplification or mutation, are frequently observed in several malignancies, most notably in non-small-cell lung carcinoma (NSCLC), breast cancer, and colorectal cancer [7,8,9,10,11,12]. The clinical effectiveness of agents designed to inhibit EGFR like gefitinib and erlotinib I [13], is significantly hindered by resistance-conferring mutations like L858R and T790M, which alter kinase domain conformation and reduce drug binding affinity. Therefore, research has shifted toward 2nd- and 3rd-generation inhibitors, such as, afatinib (2nd) and Osimertinib (3rd), effective against both wild-type and mutant EGFR isoforms. For completeness, it is important to acknowledge the EGFR 4th generation emergence inhibitors like BAY2927088, which were created especially to get beyond the reluctance of earlier generations, including the T790M and C797S mutations. These novel agents demonstrate improved selectivity and potency against resistant EGFR variants, highlighting the ongoing evolution of targeted therapies and the need for new scaffolds capable of dual or multi-target activity [14,15,16,17].
Heterocyclic scaffolds, notably pyrido[2,3-d]pyrimidines, have attracted attention due to their structural closeness to purines and varied biological characteristics [18,19,20,21,22,23,24,25,26]. These frameworks demonstrate a variety of actions, including anticancer and kinase inhibition, as well as antibacterial, antileishmanial, antitubercular, and anti-inflammatory properties [21,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52]. Several derivatives of pyrido[2,3-d]pyrimidine have shown promising EGFR inhibitory effects IIIV, as well as activity against resistant mutations VIVII [20,53,54,55,56,57] (Figure 1), making them good candidates for developing next-generation inhibitors effective against mutant EGFR in cancer drug development [58,59,60,61].
Cancer patients are frequently prone to opportunistic bacterial and viral infections caused by chemotherapy-induced immunosuppression. This co-infection scenario, particularly when antibiotic resistance is prevalent, creates significant therapeutic problems. Designing compounds with anticancer and antibacterial properties against therapeutically important pathogens including E. coli, S. aureus, E. faecalis, and P. aeruginosa is a valuable and novel technique [62,63,64,65,66,67].
According to Elsaedany et al. [68], 4-Chloro-5,7-diphenyl-2-methylpyrido[2,3-d]pyrimidine VIII shown good antibacterial efficacy against tested pathogens, particularly against Staphylococcus aureus, where it was twice as powerful as the reference. Pipemidic acid IX inhibits DNA Gyrase, making it useful for treating bacterial infections, enteritis, diarrhea, and urinary tract infections [69]. Camarasa et al. discovered that 6-(4-fluorophenyl)-2-(phenylamino)pyrido[2,3-d]pyrimidine-4,7-dione X had most potent anti-HCV (Hepatitis C virus) activity profile among all the derivatives of pyrido[2,3-d]pyrimidine tested [70]. Following testing of pyrido[2,3-d]pyrimidine-based HCV NS5A inhibitors, derivative XI was selected to be administered to a chimpanzee afflicted with HCV, and it demonstrated antiviral activity in vivo [71] (Figure 2).
Based on these findings and previous lead XII [72] (Figure 3), which has dual anticancer and antimicrobial activities, 2,4-dioxopyrido[2,3-d]pyrimidines series designed using different drug design strategies, such as keeping the main scaffold (2,4-dioxopyrido[2,3-d]pyrimidine), variation in substituent at positions 1, 3, and 7, and simplification and/or ring substitution at positions 5 and 6. These derivatives were produced by reacting the parent scaffold with different amines and assessed for cytotoxicity against the MCF-7, HCT-116, and HepG-2 cancer cell lines. Next, derivatives which were assessed as the most potent, in comparison to the EGFR wild-type and mutant isoforms (L858R and T790M). Their effects on the advancement of the MCF-7 cell cycle, apoptosis, and antibacterial activity was tested against E. coli (ATCC 25915), S. aureus (ATCC 25923), E. faecalis (ATCC 29212), and P. aeruginosa (ATCC 10145) were investigated and Finally, in silico molecular docking, ADMET, and DFT experiments validated the biological findings and provided guidance for future optimization.
To be thorough, it is necessary to mention the emergence of fourth-generation EGFR inhibitors, which have been particularly created to address resistance associated with prior generations, including the T790M and C797S mutations. These novel medicines show increased selectivity and effectiveness against resistant EGFR variants, emphasizing the continual evolution of targeted treatments and the need for new scaffolds with dual or multi-target action.

2. Results and Discussion

2.1. Chemical Part

Scheme 1, Scheme 2, Scheme 3, Scheme 4, Scheme 5, Scheme 6 and Scheme 7 outline the synthesis of targets 116. Scheme 1 shows the stepwise preparation of pyridopyrimidine analogues 15.
The key precursor, 6-amino-1,3-dimethyluracil [73], reacted with ethyl (E)-3-(4-chlorophenyl)-2-cyanoacrylate in anhydrous DMF/Et3N to afford compound 1 (ethyl amino tetrahydropyridopyrimidine-6-carboxylate). Compound 1 was subsequently treated with hydrazine hydrate, 5-hydrazinyl-1H-pyrazol-3-amine, thiocarbohydrazide, and thiosemicarbazide, producing derivatives 25.
Spectral analysis confirmed the structures. Compound 1 was fully characterized (IR, 1H NMR, 13C NMR), representing its first report. Compound 2 showed ester–carbohydrazide conversion: IR (ν NH2 at 3182 cm−1; ν C=O shift to 1625 cm−1), 1H NMR (ethyl signals lost, new NH2/NH at 6.24 and 9.10 ppm), and 13C NMR (amide carbonyl 162.87 ppm; ester signals absent).
Compound 3 confirmed ester–amide conversion and pyrazole formation: IR (ν NH2/NH at 3348, 3202 cm−1; C=N at 1517 cm−1), 1H NMR (D2O-exchangeable NH2 at 4.43 ppm; pyrazole CH at 7.82 ppm), 13C NMR (C–NHNH2 at 118.58 ppm, pyrazole CH 127.67 ppm, C=N 142.98 ppm).
Compound 4 showed incorporation of pyrazole and thiol groups with ester removal: IR (ν NH2 at 3354, 3204 cm−1; ν SH at 2346 cm−1; C=N at 1551, 1507 cm−1), 1H NMR (NH2 at 5.09 ppm; SH at 13.45 ppm), 13C NMR (pyrazole–pyridine C at 107.35 ppm; triazole C=N at 152.04, 169.00 ppm), MS (m/z 430.63 [M+], base peak 186).
Compound 5 confirmed thiosemicarbazone formation: IR (ν NH2/NH at 3349, 3254 cm−1; ν C=O at 1597 cm−1; ν C=S at 1090 cm−1), 1H NMR (exchangeable NH2/NH at 3.01, 10.89, 12.06 ppm; ethyl signals absent), 13C NMR (C=S carbonyl 186.05 ppm; amide carbonyl 171.62 ppm).
Scheme 2 depicts the synthesis of pyrazole and oxadiazole derivatives 68 from Compound 2. Reaction with acetylacetone in ethanolic sodium ethoxide yielded dimethyl-substituted pyrazole 6. Mercapto-oxadiazole 7 was obtained by treating Compound 2 with carbon disulfide under basic conditions, while cyclized derivative 8 was formed via reflux with malononitrile. All structures were confirmed by spectral analysis.
Spectral data confirmed the structures of Compounds 68. For 6, IR showed ν CH3 at 2819/2736 cm−1, ν C=O shifted 1625 → 1740 cm−1, and ν C=N at 1530 cm−1, with NH2/NH bands absent. 1H NMR displayed CH3 singlets at 2.71/2.87 ppm and pyrazole CH at 6.24 ppm; 13C NMR showed CH3 at 13.04/13.97 ppm, pyrazole CH at 110.02 ppm, C–CH3 at 141.64 ppm, and C=N at 151.06 ppm. MS gave m/z 438.87 [M+], base peak 132.
Compound 7’s mercapto-oxadiazole formation was confirmed by IR (ν C=N 1599/1588 cm−1, ν SH 2596 cm−1; NH2/NH/amide absent), 1H NMR SH singlet at 13.80 ppm, and 13C NMR pyridine–oxadiazole at 107.77 ppm, C=N at 166.66/169.59 ppm; amide carbonyl disappeared.
For 8, IR showed NH2 at 3416/3216 cm−1, cyano at 2222 cm−1, pyrazole C=N at 1555/1517 cm−1, with NH and amide C=O absent. 1H NMR displayed pyrazole CH at 2.79 ppm and NH2 at 6.23 ppm (D2O exchangeable); 13C NMR showed pyrazole C at 33.50 ppm, pyridine C at 100.24 ppm, CN at 118.03 ppm, and C=N at 165.80/186.90 ppm; amide carbonyl at 166.61 ppm disappeared.
In Scheme 3, the oxadiazol propanenitrile derivative 9 was prepared by refluxing compound 7 with acrylonitrile in a basic media, followed by treatment with sodium azid and ammonium hydrochloride to produce tetrazole derivative 10.
Spectral data confirmed the structures of Compounds 9 and 10. For 9, IR showed CH2 stretches at 2989/2873 cm−1, ν C≡N at 2249 cm−1, C=S at 1058 cm−1, and disappearance of ν SH (2596 cm−1). 1H NMR displayed acrylonitrile CH2 triplets at 3.91/4.15 ppm, with SH proton absent at 13.80 ppm. 13C NMR showed CH2 at 16.10/45.70 ppm, CN at 118.03 ppm, and C=S at 174.00 ppm, while the previous oxadiazole C=N at 169.59 ppm disappeared.
Compound 10 formation of the tetrazole ring was confirmed by IR (ν NH 3282 cm−1, ν C=N 1491 cm−1, ν N=N 1297 cm−1; disappearance of ν C≡N 2249 cm−1), 1H NMR (CH2 shifted 3.91 → 2.71 ppm; D2O-exchangeable NH singlet at 14.00 ppm), and 13C NMR (CH2 at 22.00/54.89 ppm, C=N at 165.94 ppm; CN at 118.03 ppm absent). MS showed m/z 512.93 [M+], base peak 80.
Scheme 4 depicts the planned molecular method for synthesizing the novel fused pyrazole pyrimidinone and amino pyrimidine derivatives 11 and 12.
Compound 8 was fused with formic acid and formamide under sand bath heating to yield 5,6-dihydropyrimidin-4(3H)-one (11) and 5,6-dihydropyrimidin-4-amine (12).
For 11, IR showed ν NH 3161 cm−1, ν CH 2870 cm−1, lactam C=O 1723 cm−1, and pyrimidinone C=N 1635 cm−1, with disappearance of NH2 and cyano bands (3416, 3216, 2222 cm−1). 1H NMR displayed pyrazole CH at 2.32 ppm, pyrimidinone CH at 8.38 ppm, and NH at 11.44 ppm (D2O-exchangeable). 13C NMR showed pyrazole CH 46.83 ppm, pyrimidinone C=N 150.98 ppm, lactam C=O 169.85 ppm, pyrazole C=N 181.01 ppm; cyano carbon at 118.03 ppm disappeared. MS gave m/z 450.84 [M+], base peak 319.
For 12, IR revealed NH2 at 3422 cm−1 and pyrimidin-4-amine C=N at 1610/1594 cm−1; cyano band at 2222 cm−1 disappeared. 1H NMR showed pyrazole CH at 1.65 ppm, D2O-exchangeable NH2 at 8.25 ppm, and pyrimidine CH at 8.91 ppm. 13C NMR displayed pyrazole CH 35.86 ppm, pyrimidine C=N 162.42/163.09 ppm, pyrazole C=N 167.12/178.02 ppm; cyano carbon absent, confirming formation.
The dimethyl pyrazole derivative 13 was effectively produced by refluxing compound 3 with acetylacetone in a basic solution while employing triethylamine as a catalyst (Scheme 5).
Spectral data confirmed Compound 13 as a dimethyl-substituted pyrazole. IR showed ν C=N at 1491 cm−1. 1H NMR displayed CH3 singlets at 1.89/2.10 ppm and pyrazole CH at 5.04 ppm; no D2O-exchangeable NH2/NH signals were observed. 13C NMR showed CH3 at 12.32/13.50 ppm, pyrazole C–CH3 90.00 ppm, CH 112.25 ppm, C=N 142.20 ppm, and quaternary C 155.35 ppm. MS gave m/z 519 [M+], base peak 300.
In Scheme 6, product 4 was exposed to further reactions with 2-chloroacetyl chloride and carbon disulfide, yielding the corresponding derivatives: 5-chloro-thiadiazine (compound 14) and 1,3-thiadiazole-2-thiol (compound 15).
Spectral data confirmed Compounds 14 and 15. For 14, IR showed ν C=N at 1491 cm−1 and disappearance of ν SH (2346 cm−1), 1H NMR displayed CH2 singlet at 4.92 ppm with SH absent, and 13C NMR showed CH2 at 34.16 ppm, C=N at 151.06 ppm, and imine triazole C shifted 169.00 → 162.80 ppm.
Compound 15’s 1,3-thiadiazole-2-thiol formation was confirmed by IR (ν SH 2550 cm−1), 1H NMR (SH singlet at 13.24 ppm), 13C NMR (imine C 188.17 ppm), and MS m/z 572 [M+], base peak 284.
Finally, Scheme 7 produced 5-methyl pyrazolone derivative 16 by refluxing compound 5 with ethyl acetoacetate in a strongly basic solution containing sodium hydride as the base.
Spectral data confirmed Compound 16 as a 5-methyl pyrazolone. IR showed ν CH 2839/2816 cm−1, C=O 1727 cm−1, and C=N 1516 cm−1. 1H NMR displayed CH3 singlet at 1.94 ppm and CH2 at 3.01 ppm. 13C NMR showed CH3 16.00 ppm, CH2 45.05 ppm, C=N 159.58 ppm, C=O 165.33 ppm, and C=S shifted 186.05 → 182.25 ppm. MS gave m/z 499 [M+], base peak 245.

2.2. Biological Evaluation

2.2.1. Antiproliferative In Vitro Potency

Cytotoxicity testing of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives (116) on HeLa, HepG-2, and MCF-7 cells via MTT assay [74,75,76] showed variable micromolar IC50 values, comparable to erlotinib. (Figure 4, Table S2).
Compounds 1, 2, and 7 were the most active in the series, their IC50 values varying from 3.98 to 17.52 µM across all examined cell lines. Compound 1 inhibited MCF-7 cells the greatest (IC50 = 3.98 ± 0.2 µM), outperforming erlotinib (IC50 = 7.26 ± 0.3 µM) in this line. It also had considerable activity against HeLa and HepG-2 cells. Furthermore, compound 7 showed significant cytotoxicity, especially against HeLa (IC50 = 9.72 ± 0.9 µM). In contrast, derivatives 5, 8, 9, 10, 11, and 14 showed modest action, with IC50 values >50 µM in most cases, indicating lower affinity for the molecular target(s) or poor cell permeability.
In addition, all derivatives were tested in vitro towards normal lung fibroblast WI-38 cell lines to determine their safety. The ideal IC50 values for potential derivatives 1 and 2 are 81.65 ± 4.1 and 64.07 ± 3.5 µM, respectively, while erlotinib has an IC50 of 78.32 ± 3.9 µM. Consequently, they may be classified as anticancer drugs with acceptable safety. A chemical’s selectivity and capacity to target cancer cells without endangering healthy cells are gauged by its SI value (WI-38); the greater the value, the safer the substance. The selectivity index (SI) toward tumor cells was measured as the IC50 value in WI-38 fibroblasts relative to that in malignant cell lines.
Best Compounds Based on SI (Selectivity)
  • Compound 1 achieved the highest SI values, making it the most selective among all compounds: 20.51 on MCF-7 cells, 13.83 on HepG-2 cells, and 12.98 on HeLa cells.
  • Compound 2 also showed good selectivity, but it was lower than both Compound 1 and Erlotinib. Based on this analysis, Future studies should concentrate on comprehending Compound 1’s mode of action and how it achieves this exceptional efficacy and selectivity.
Structure-Activity Relationship Study
The cytotoxic evaluation of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives towards cell lines (HeLa, HepG-2, and MCF-7) demonstrated a clear influence of C-6 substitution type, size, and electronic character on anticancer activity. Compound 1, ethyl 7-amino-5-(4-chlorophenyl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine-6-carboxylate, containing a C-6 ethyl ester, displayed the highest efficacy (IC50: 3.98–6.29 µM), exceeding erlotinib in MCF-7 cells. The ethyl ester’s modest bulk and lipophilicity are anticipated to increase hydrophobic interactions in the binding pocket while retaining high cell permeability.
Replacing the ester with the more polar 6-carbohydrazide in 2 decreased activity (IC50: 7.54–12.94 µM). The enhanced polarity and hydrogen-bonding ability may boost solubility while disrupting hydrophobic fit. Adding a bulkier N-(4-hydrazinyl-1H-pyrazol-3-yl)-6-carboxamide in 3 or 6-(4-amino-5-mercapto-4H-1,2,4-triazol-3-yl) in 4 reduced potency (14–26 µM), perhaps due to steric interference with optimum binding and altered electron distribution at the C-6 region.
The N-(hydrazinecarbonothioyl)-6-carboxamide in 5 resulted in a significant loss of activity (IC50 > 84 µM), probably due to extra steric bulk and high electron density around the thioamide, which may repel electron-rich protein residues or hamper structural fit. The 6-(3,5-dimethyl-1H-pyrazole-1-carbonyl) substitution in 6 maintained considerable efficacy (24–32 µM), with the planar heterocycle helping π-π interactions, but was lowered compared to the ester due to higher steric hindrance from the dimethyl groups.
In contrast, 6-(5-mercapto-1,3,4-oxadiazol-2-yl) in 7 exhibited significantly higher potency (IC50: 9.72–17.52 µM), comparable to compounds 1 and 2. The oxadiazole ring has electron-drawing properties and a compact shape, promoting strong interactions with little steric penalty. Adding 3-amino-4H-pyrazole-4-carbonitrile in 8 or 2-thioxo-1,3,4-oxadiazol-3(2H)-yl propanenitrile in 9 significantly decreased potency (62–88 µM). The nitrile groups are significantly electron-withdrawing, potentially disrupting crucial hydrogen-bond donor-acceptor balances, and the bulky appendages are expected to produce steric conflict. The tetrazolyl-oxadiazole-thione 10 and fused pyrazolo[3,4-d]pyrimidine derivatives 11, 12 had low activity (44–73 µM).
These rigid, bulky complexes are anticipated to limit conformational flexibility in the binding site and modify electron distribution in a way that lowers affinity. The N-(3,5-dimethyl-1′H-[1,4′-bipyrazol]-3′-yl)-6-carboxamide 13 showed low-moderate potency (28–36 µM), whereas 6-(6-chloro-7H-[1,2,4]triazolo [3,4-b][1,3,4]thiadiazin-3-yl) 14 was among the weakest (69–94 µM). In the latter, bulk and chlorine substitutions may cause steric and electrical mismatches with the binding site. The efficacy of 6-(6-mercapto-[1,2,4]triazolo [3,4-b][1,3,4]thiadiazol-3-yl) 15 increased slightly over 14 (40–52 µM) due to possible extra thiol-mediated interactions, but remained low. N-(3-methyl-5-oxo-4,5-dihydro-1H-pyrazole-1-carbonothioyl)-6-carboxamide 16 had modest activity (35–65 µM), possibly due to steric and electron-rich sulfur effects that disrupted ideal orientation.
In conclusion, the best activity requires moderate-sized, slightly lipophilic, and electron-balanced groups (e.g., ethyl ester in 1, carbohydrazide in 2, and mercapto-oxadiazole in 7). Loss of activity is associated with high polarity (hydrazide), excessive steric bulk (polycyclic, fused rings), or strongly electron-rich sulfur-rich thioamides. Electron-withdrawing substituents can be beneficial when combined with compact geometry (oxadiazole in 7) but detrimental when paired with bulky groups (nitrile-extended heterocycles 8 and 9). Finally, planarity and reduced steric hindrance at C-6 appear essential for maintaining strong cytotoxic activity across cancer cell lines (Figure 5).

2.2.2. In Vitro Inhibitory Activity Evaluation Against Wild EGFR, Mutant EGFR (L858R), and EGFR (T790M)

The synthesized 1,3-dimethyl uracil derivatives 1 and 2 inhibitory potency was tested in vitro against wild-type EGFR, the sensitizing L858R mutant, and the resistance-associated T790M mutant, and compared to the clinically authorized EGFR-TKI erlotinib [77,78] (Table 1).
The IC50 of compound 1 was 0.093 ± 0.003 μM towards wild-type EGFR, comparable to erlotinib (0.051 ± 0.002 μM) and superior to compound 2 (0.212 ± 0.007 μM). This high activity indicates that the planar heteroaromatic pyridopyrimidine scaffold effectively engages the ATP-binding pocket, establishing critical hinge area interactions similar to those found in erlotinib.
For the L858R mutant, a clinically important activating mutation that boosts EGFR kinase activity, both drugs retained significant activity although at a lower potency than wild-type EGFR. Compound 1 (1.157 ± 0.036 μM) had a 12-fold lower potency than wild-type EGFR, while compound 2 (1.603 ± 0.050 μM) showed a 7.5-fold drop. Erlotinib (0.239 ± 0.007 μM) resulted in a smaller ~4.7-fold reduction. This lower efficacy in the L858R version could be attributed to small conformational changes within the kinase domain that affect inhibitor accommodation, a tendency shared by other ATP-competitive TKIs.
Compound 1 exhibited substantial activity (0.174 ± 0.006 μM) against the T790M gatekeeper mutant, which causes resistance to first-generation EGFR inhibitors. It was only about 2-fold less powerful than against wild-type EGFR. In comparison, erlotinib (0.098 ± 0.003 μM) exhibited a comparable potency drop, while compound 2 (0.854 ± 0.027 μM) showed a more significant activity loss (~4-fold reduction). The significant T790M activity of compound 1 indicates that the scaffold can partially overcome the steric hindrance created by the methionine substitution, possibly because to favorable hydrophobic and hydrogen bonding interactions within the changed ATP pocket.

2.2.3. 1,3-Dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1 Cell Cycle Arrest and Apoptosis

Cell cycle and apoptosis assays using flow cytometry and annexin-V staining [79,80] were conducted to explore the mechanism behind the cytotoxic effect of compound 1, the most active derivative of the tested 1,3-dimethyl-2,4-dioxopyrido [2,3-d]pyrimidine series.
Flow cytometric analysis after 48 h of treatment with compound 1 (Table S4, Figure 6 and Figure 7) showed that cells of MCF-7 were accumulated significantly in the G0/G1 phase (74.86%) compared to untreated controls (56.29%). Alongside this, the S phase population significantly decreased (22.79% vs. 35.72% in control), and a sharp decline in the G2/M phase fraction (1.75% vs. 7.99% in control). These results suggest that compound 1 triggers G1 phase cell cycle arrest, effectively halting cell cycle progression before DNA synthesis.
Apoptosis analysis (Table S5, Figure 6 and Figure 8) demonstrated a marked increase in total apoptotic cells upon exposure to compound 1 (26.37%) compared to control cells (3.18%). The apoptotic population comprised both early apoptosis (5.86%) and a predominant late apoptosis fraction (16.62%), with a modest increase in necrosis (3.89% vs. 2.31% in control). The dominance of late apoptosis suggests that compound 1 not only triggers apoptotic signaling but also drives cells to complete programmed cell death within the 48 h treatment period. The significant induction of apoptosis, together with G0/G1 arrest, points to a dual mechanism whereby compound 1 suppresses proliferation and promotes elimination of cancer cells through apoptotic pathways.

2.2.4. Wound Healing Assay

The preformation of wound healing (scratch) assay to assess the migration ability of MCF-7 cells following treatment with 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1, compared to untreated controls. After 72 h, the control group exhibited 94.82 ± 3.50% wound closure, indicating near-complete gap filling, while compound 1 treatment achieved 69.63 ± 2.57% closure. This reduction was supported by shorter migration lengths (0.31–0.32 mm vs. 0.42–0.43 mm in controls) and smaller migrated areas (0.564 mm2 vs. 0.768 mm2). The marked decrease in closure percentage suggests that derivative 1 impairs cellular migration in MCF-7 cells (Figure 9, Table S6).
Overall, the 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1’s ability to limit wound closure suggests potential as an anti-metastatic agent in breast cancer therapy.

2.2.5. In Vitro Antibacterial Activity

Antibacterial activity evaluation was performed using four bacterial strain types: Staphylococcus aureus (ATCC25923), Escherichia coli (ATCC25915), Enterococcus faecalis (ATCC29212), and Pseudomonas aeruginosa (ATCC10145) by the well diffusion method [81,82,83,84] (Table S7, Figure 10).
Table 2 presents the values of MIC (µg/mL) for compounds 1, 12, 13, and 16 against four bacterial strains: E. coli ATCC25915, S. aureus ATCC25923, E. faecalis ATCC29212, and P. aeruginosa ATCC10145.
Compound 1 has moderate activity, with MICs of 50 µg/mL against E. coli and P. aeruginosa and 100 µg/mL against S. aureus and E. faecalis, demonstrating balanced efficacy across strains. Compound 12 has similar action against E. coli and P. aeruginosa (50 µg/mL), but less effective against S. aureus (>100 µg/mL) and E. faecalis (100 µg/mL), indicating selective potency.
Compound 13 consistently inhibits E. coli at 50 µg/mL but shows reduced effectiveness against S. aureus (>100 µg/mL), E. faecalis (>100 µg/mL), and P. aeruginosa (100 µg/mL), hinting at lower activity against Gram-positive bacteria. Similarly, Compound 16 matches the 50 µg/mL MIC against E. coli and P. aeruginosa, with moderate efficacy at 100 µg/mL against S. aureus and E. faecalis. Overall, the compounds demonstrate the lowest MICs against E. coli and P. aeruginosa, indicating better activity against these Gram-negative bacteria, while S. aureus appears least susceptible with MICs ≥ 100 µg/mL [85].
Although the antibacterial activity was moderate compared to specialized standard antibiotics ciprofloxacin, the demonstrated effect against resilient Gram-negative pathogens highlights the scaffold pyrido[2,3-d]pyrimidine potential for further development as a multi-target therapeutic agent.

2.3. Computational Studies

2.3.1. ADMET In Silico Prediction

ADMET computational profiling by free online resources SwissADME and AdmetSAR 1.0 [86,87,88] demonstrated that 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 1 and 2 satisfy Lipinski’s Rule of Five and Veber’s criteria, with molecular weights < 500 Da, optimal TPSA, and MLogP values, supporting oral drug-likeness (Table 3, Figure 11A). Both analogues showed high predicted human intestinal absorption and were non-substrates of P-glycoprotein, reducing the likelihood of efflux-related bioavailability loss. The absence of predicted blood–brain barrier permeability suggests suitability for non-CNS indications (Figure 11B).
Distribution projections revealed organelle-specific localization, with compound 1 preferring lysosomes and compound 2 targeting mitochondria, which might be used for targeted therapeutic action. Metabolic study revealed that both are non-inhibitors of the majority of CYPs, implying a low frequency of drug–drug interactions and little effect on other active substances. Toxicity predictions revealed weak hERG I inhibition in both compounds, necessitating cardiotoxicity testing, although Ames and carcinogenicity tests were negative. Both have acute oral toxicity class III (771.1 and 728.4 mg·kg−1, respectively) and are not easily biodegradable, creating environmental persistence concerns (Table S8). Overall, the ADMET profile encourages further research into both drugs as orally active options with tolerable hazards.

2.3.2. Docking Simulation

To justify the in vitro cytotoxic and enzyme inhibitory actions of Compounds 1 and 2, the most potent synthetic 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives, against important cancer-related targets, EGFRWT and mutant EGFRT790M, molecular docking studies were conducted. Docking simulations were carried out with MOE-Dock (version 2024.0601) [89,90,91,92]. The crystal structures of EGFRWT and EGFRT790M, complexed with their reference ligands erlotinib and WZ4002 (PDB IDs: 1M17 and 3IKA) [78,93], were retrieved from the Protein Data Bank. Validation of the docking protocol using these native ligands resulted in low RMSD values (0.91 and 1.27 Å) and favorable binding energies (−11.14 and −10.31 kcal/mol), confirming the reliability of the procedure.
Figure 12 shows that the derivatives 1 and 2 were positioned correctly within EGFR showing favorable binding energies of −10.93 and −10.68 kcal/mol, respectively. The oxygen of dioxopyrido[2,3-d]pyrimidine scaffold at p-2, afforded H-bond acceptor with the Met769 backbone in both derivatives, resembling the native ligand, and additional one with Leu768 in 1. Also, the methyl group at p-1 in compound 1 displayed an H-bond donor with Gln767 backbone (distance: 3.32 Å). The sidechain of Asp831 accepted one hydrogen bond from the amino group at p-7 in 1 and two from the hydrazide fragment at p-6 in 2 (distances: 2.77, 2.50, and 2.88 Å, respectively). Also, the centroid of 4-chlorophenyl in 1 applied arene-H interaction with Lys721.
By inspection of the EGFRT790M receptor in Figure 13, the analogues 1 and 2 gave comparable energy scores of –9.91 and –9.22 kcal/mol, respectively. Also, H-bond acceptors were detected between the oxygen of dioxopyrido[2,3-d]pyrimidine scaffold at p-2 and the Met793 backbone in both derivatives, and another one with Leu792 in 1 (distances: 2.47, 3.11, and 3.08 Å, respectively). At p-6 of dioxopyrido[2,3-d]pyrimidine scaffold, an H-bond donor was demonstrated between the carboxylate in 1 and another one between hydrazide in 2 and Asn842 (distances: 3.21 and 2.60 Å, respectively).
To confirm the mechanism of the predicted antimicrobial and antiviral potency of our target 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 1, a docking study was applied using the receptors of E. coli DNA gyrase and HCV NS5A (PDB codes: 1AJ6 & 3FQM, respectively) [94,95]. After validation of both receptors, giving tiny RMSD values of 0.78 and 1.38 Å and energy scores of −10.58 and −9.82 kcal/mol, respectively. Docking of compound 1 with both active sites resulted in promising energy scores of −10.25 and −9.36 kcal/mol, respectively, indicating perfect binding. Three H-bonds were created between the carbonyl oxygen at p-4, carboxylate at p-6, and the amino group at p-7 with the key amino acids Asp46, Asp73, and Glu50 sidechains (distances: 2.74, 3.33, and 3.29 Å, respectively) (Figure 14).
Regarding HCV NS5A ATP-pocket, the carbonyl oxygen of dioxopyrido[2,3-d]pyrimidine scaffold at p-2 in 1 exhibited three H-bond acceptors, one with Ala92 backbone and two with Arg157 sidechain (distances: 3.17, 2.79, and 3.07 Å, respectively) (Figure 15). Furthermore, the centroids of pyrimidine and 4-chlorophenyl revealed arene-H interactions with Arg157 and Tyr161, respectively.

2.3.3. Molecular Dynamics Simulations

Molecular dynamics simulations were subsequently applied to confirm the docking outcomes of compound 1 (1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine) within the binding domains of both EGFRWT and the T790M mutant [77,96]. The results, illustrated in Figure 16, Figure 17, Figure 18 and Figure 19, summarize the dynamic trajectories, including protein–ligand interactions, RMSD, and RMSF profiles.
Both the free EGFRWT and mutant EGFRT790M enzymes were found to be highly dynamic by RMSD estimates. The unbound protein of EGFRWT achieved an average RMSD value of 9 Å in Figure 16A, which decreased upon binding with 1 (3 Å). Also, the plot of RMSD of 1 within EGFRT790M indicated that the complex reached stability at 30 ns within the range of 3.5–4.6 Å, throughout the simulation (Figure 16B), suggesting that compound 1 was highly successful in reducing the dynamic behavior of EGFRWT, and mutant EGFRT790M.
With ranges of 1.8–3 Å and 0.9–2.0 Å for 1 within EGFRWT, and mutant EGFRT790M, respectively, the RMSF plots of protein-compound 1 complexes (Figure 17A,B) reflected a notable reduction. This suggests that derivative 1 remains tightly accommodated in the binding cavities of both EGFRWT and the T790M mutant, thereby contributing to the stability of the formed complexes.
Derivative 1 continuously maintained a hydrogen connection with Met769 and Asp831 for 100% and 40% of the simulation time within EGFRWT, respectively (Figure 18A), demonstrating strong binding. However, within EGFRT790M, H-bonding with the residues Glu762, Met793, and Asp855 was highly observed for 40%, 98% and 43% of the simulation time, respectively (Figure 18B). Additionally, an intense orange color interacting with residues Phe699, Leu768, Met769, Leu820, and Asp831 within EGFRWT (Figure 19A) and with Phe723, Glu762, Met793, Leu844, and Asp855 within EGFRT790M revealed a remarkable situation of interactions for the screened derivative 1 with EGFRWT, and mutant EGFRT790M residues over 100ns (Figure 19A,B).

2.3.4. Quantum Chemical Calculation

Computational Profiling and Global Reactivity Descriptors
A series of sixteen compounds was evaluated using quantum chemical descriptors, including HOMO, LUMO, and the energy gap (ΔE), as well as conceptual DFT parameters: electronegativity (χ), ionization potential (I), chemical hardness (η), softness (S), electrophilicity index (ω), and charge transfer indices (ΔN and ΔN_max). Significant variations in the HOMO–LUMO gap were observed across the series. Compound 2 exhibited the narrowest gap (ΔE = 0.00478 eV), corresponding to extremely high softness (S = 418.41) and electrophilicity (ω = 12.89). In contrast, compound 1 had the widest gap (ΔE = 0.1064 eV) with lower softness (S = 18.80) and reduced electrophilicity (ω ≈ 0.92). The detailed quantum chemical values are summarized in Table 4, and the DFT-optimized structures along with HOMO/LUMO diagrams for selected compounds are provided in the Supplementary Information (Table S9). These results highlight the diverse electronic behavior of the compounds, suggesting potential structure–activity relationships.
Frontier Molecular Orbitals and Hardness/Softness
Compound 1 has a very negative HOMO value of −0.2744, indicating a reduced tendency to donate electrons. Compounds 1, 7, and 9 have greater negative LUMO values of −0.16, indicating a higher ability to take electrons. According to the HSAB principle, molecules with smaller ΔE values are softer, more chemically reactive, and more prone to engage in polar interactions with biological targets. However, excessive softness, as observed in compound 2, can result in uncontrolled reactivity and poor selectivity, emphasizing the importance of balanced electrical characteristics. geometries along with the HOMO and LUMO orbitals for the two top-ranked compounds (1 and 2) in Figure 20.
Electrostatic Potential (ESP) Mapping and Areal Distribution
Molecular electrostatic potential (ESP) surface analysis provided insights into the distribution of electron density. Positive (blue/cyan), negative (red/orange/yellow), and neutral (green) surface areas were quantified. ESP areas were extracted from color-thresholded MEP images using Python (OpenCV 3.13.7), and the resulting percentages were applied in the calculation of area balance and incorporated into the ESP Score.
Representative examples include compound 5 (Positive 26.86%, Negative 8.80%, area balance ≈ +18.06) and compound 12 (Positive 24.89%, Negative 6.00%, area balance ≈ +18.89), both showing strongly positive ESP distributions. Conversely, compound 2 displayed a predominantly negative ESP distribution (Positive 9.85%, Negative 14.82%, balance ≈ −4.97). These findings indicate that both global descriptors and localized ESP features are crucial in predicting bioactivity, as extended positive regions may enhance interactions with nucleophilic residues at biological target sites (see Supplementary Information, Tables S10 and S11).
Composite Bioactivity Score and Compound Ranking
A composite bioactivity score was calculated as: Composite Bioactivity Score = 0.30 × n(Softness) + 0.20 × n(ω) + 0.20 × n(inv ΔE) + 0.15 × ESP Score, where the ESP Score was derived from normalized ESP maxima, minima, contrast, and area balance.
Based on this model, the top-ranked compounds were:
  • Compound 5 (1.0834): Balanced softness (S ≈ 34.30), moderate electrophilicity (ω ≈ 1.35), favorable ESP profile.
  • Compound 12 (1.0668): Despite a higher ΔE (0.0776 eV), advantageous ESP distribution enhanced its ranking.
  • Compound 2 (0.8306): Extremely low ΔE and high softness/electrophilicity, but negative ESP bias reduced its score.
  • Compounds 4 (0.6380) and 1 (0.6363): Moderate predicted activity.
Intermediate activity was predicted for compounds 16 (0.3367), 13 (0.3264), 11 (0.2907), 15 (0.2756), 9 (0.2611), and 8 (0.2447), while compounds 7 (0.1758) and 10 (0.1695) were least favorable (Table 5).
Interplay Between Global and Local Descriptors
While softness and electrophilicity strongly influence predicted activity, ESP surface profiles were decisive in differentiating top-ranked compounds. Compounds 5 and 12 benefited from extended positive ESP regions and favorable area balances, compensating for moderate ΔE values. Conversely, compound 2, despite remarkable global reactivity, was penalized by its negatively skewed ESP distribution.
Charge Transfer Capacity
High ΔN max values, as observed for compound 2, indicate enhanced electron-accepting potential, consistent with its high ω and S values. However, this alone was insufficient for top ranking without complementary ESP features, highlighting the importance of integrating both global and local descriptors for reliable activity prediction.
Implications for SAR and Lead Prioritization
Compounds 5, 12, and 2 were identified as the top candidates for subsequent biological studies. Structural modifications that enhance positive ESP regions while maintaining moderate ΔE values are likely to improve activity. Additionally, avoiding excessive softness while preserving favorable ESP contrast and balance is recommended to ensure stability and selectivity.
Limitations and Outlook
The scoring system relies on internal min–max normalization, and results may vary with other chemical datasets. ESP features were derived from static conformations, which may change in different phases or conformers. Nevertheless, the agreement between compound rankings, conceptual DFT principles, and ESP distributions supports the model’s robustness. Experimental validation via docking, molecular dynamics, and ADMET studies is essential to confirm predictions.
Target Compound Selection (Compound 1)
Although compounds 5 and 12 scored highest, compound 1 was selected as the target molecule due to its balanced electronic properties and superior stability. With the widest HOMO–LUMO gap (ΔE = 0.1064 eV), moderate softness (S ≈ 18.80), and relatively low electrophilicity (ω ≈ 0.92), compound 1 is expected to interact selectively with biological targets, combining synthetic feasibility, electronic stability, and acceptable reactivity. Therefore, it was advanced as the lead for subsequent experimental validation and biological assays.

3. Materials and Methods

3.1. Chemistry

The instruments employed included those for determining melting points and recording IR, 1H NMR, 13C NMR, and mass spectra are described in detail in the supplemental material file.
Using the previously published method, 6-amino-1,3-dimethylpyrimidine-2,4(1H,3H)-dione, the starting chemical, was produced [73].

3.1.1. Synthesis of Ethyl 7-amino-5-(4-chlorophenyl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine-6-carboxylate (1)

A mixture of 6-amino-1,3-dimethylpyrimidine-2,4(1H,3H)-dione (10 mmol, 1.55 g) and ethyl (E)-3-(4-chlorophenyl)-2-cyanoacrylate (10 mmol, 2.35 g) in 10 mL of anhydrous DMF with a catalytic quantity of triethylamine present. For 6h, the reaction was refluxed at 80–100 °C. Using eluent ethyl acetate: hexane (1:1), TLC was performed to track the progress of the reaction. Upon completion, the reaction mixture was cooled to room temperature and then poured into 50 mL of chilled water, resulting in precipitation. Compound 1 was obtained as a brown powder by vacuum filtering the resultant solid, extensively washed by cold ethanol, with the final step involving recrystallization from ethanol.
Yield 89%; m.p. 229–231 °C; IR (νmax/cm−1): 3410, 3332 (NH2), 3046 (CH-arom.), 2954, 2923, 2853 (CH-aliph.), 1735 (C=O ester), 1692, 1650 (2C=O amide), 1594 (C=N), 1565, 1528 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz): δ 1.71 (t, J = 7.2 Hz, 3H, CH3 ester), 3.07, 3.12 (s, 6H, 2CH3), 4.05 (q, J = 7.1 Hz, 2H, CH2), 7.30–7.89 (m, 4H, Ar-H), 8.48 (s, 2H, NH2, D2O exchangeable) ppm.; 13C-NMR (DMSO-d6, 100 MHz); δ 14.92 (1C, CH3 ester), 23.06, 26.03 (2C,2CH3), 57.80 (1C, CH2), 96.09 (1C, C pyridine-C=O ester), 102.97 (1C, C pyridine-C=O amide), 128.74, 129.37, 129.92, 130.48, 130.84, 131.39 (6C, C-Ar), 152.57 (1C, C pyridine-Ph-Cl), 155.00, 158.47 (2C, C=O uracil), 155.90 (1C, C=N pyridine), 161.97 (1C, C pyridine-N-CH3), 166.23(1C, C=O ester) ppm. Anal. Calcd for C18H17ClN4O4 (388.81): C, 55.61; H, 4.41; Cl, 9.12; N, 14.41; O, 16.46%. Found: C, 55.65; H, 4.40; Cl, 9.12; N, 14.43; O, 16.65 %.

3.1.2. 7-Amino-5-(4-chlorophenyl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine 6-carbohydrazide (2)

The solution was made by dissolving 1 (3.88 g, 10 mmol) in 10 mL of DMF while maintaining stirring at ambient temperature for 30 min. Hydrazine hydrate (1.25 mL, 25 mmol) was added gradually while stirring constantly following pre-equilibration. After that, the mixture was refluxed for 4 h at 70 °C. Using ethyl acetate: hexane (1:1) as the eluent, TLC verified the end of the reaction process. The resultant precipitate was recovered by vacuum filtering and properly cleaned with cold ethanol after cooling to RT. Compound 2, a pale yellow powder, was obtained by further purifying the crude product by recrystallizing it from methanol.
Yield 81%; m.p. 200–202 °C; IR (νmax/cm−1): 3475, 3373 (NH2-pyridne), 3400, 3256 (NH2-NH), 3182 (NH), 3046 (ν CH-arom.), 2989, 2945 (ν CH-Aliph.), 1690, 1651 (2C=O amide), 1625 (C=O amide-NH), 1590 (C=N), 1561, 1543 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 3.07, 3.12 (s, 6H, 2CH3), 6.24 (s, 2H, NH2-NH, D2O exchangeable), 7.31–8.27 (m, 4H, Ar-H), 8.47 (s, 2H, NH2-pyridine, D2O exchangeable), 9.10 (s, 1H, NH, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz); δ 23.05, 26.06 (2C, 2CH3), 102.14 (2C, 2C pyridine-C=O), 127.60, 128.42, 128.71, 129.18, 129.64, 131.37 (6C, C-Ar), 152.89 (1C, C pyridine-Ph-Cl), 154.46, 158.46 (2C, C=O uracil), 156.89 (1C, C=N pyridine), 162.14 (1C, C pyridine-N-CH3), 166.61 (1C, C=O-NH amide) ppm. Anal. Calcd for C16H15ClN6O3 (374.79): C, 51.28; H, 4.03; Cl, 9.46; N, 22.42; O, 12.81%. Found: C, 51.29; H, 4.01; Cl, 9.48; N, 22.42; O, 12.81%.

3.1.3. 7-Amino-5-(4-chlorophenyl)-N-(4-hydrazinyl-1H-pyrazol-3-yl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine-6-carboxamide (3)

A weighed portion of compound 1 (3.88 g, 10 mmol) was first dissolved in 10 mL of GL AcOH and stirred at ambient temperature for 30 min. Once equilibrium was established, 3-amino-5-hydrazinyl pyrazole (1.13 g, 10 mmol) was introduced gradually with continuous stirring. The reaction mixture was then subjected to reflux at 80 °C for eight hours. Progress of the reaction was monitored by TLC using ethyl acetate/hexane (1:1) as the mobile phase. After completion, the mixture was cooled to room temperature, and the formed solid was separated by vacuum filtration and washed thoroughly with cold ethanol. The crude material was recrystallized from ethyl acetate to yield compound 3 as a brown crystalline powder.
Yield 72%; m.p. 208–210 °C; IR (νmax/cm−1): 3348, 3202 (2NH2, 3NH), 3050 (CH-arom.), 2945 (CH-Alipha.), 1710, 1658 (2C=O amide), 1625 (C=O amide-NH), 1560 (C=N pyridine), 1517 (C=N pyrazole), 1492 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 3.04, 3.19 (s, 6H, 2CH3), 4.43 (s, 2H, NH2-NH, D2O exchangeable), 7.32–7.46 (m, 4H, Ar-H), 7.82 (s, 1H, CH pyrazole), 8.40 (s, 2H, NH2-pyridine, D2O exchangeable), 9.69 (s, 1H, NH-NH2, D2O exchangeable), 11.30 (s, 1H, NH-C=O, D2O exchangeable), 12.27 (s, 1H, NH pyrazole, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz) δ = 23.03, 26.52 (2C, 2CH3), 102.00 (2C, 2C pyridine-C=O), 118.58 (1C, C pyrazole-NHNH2), 127.67 (1C, CH pyrazole), 128.42, 128.77, 128.94, , 129.51, 130.42, 131.18 (6C, C-Ar), 142.98 (1C, C=N pyrazole), 151.13 (1C, C pyridine-Ph-Cl), 154.20, 158.39 (2C, 2C=O uracil), 156.55 (1C, C pyridine-N-CH3), 158.99 (1C, C=N pyridine), 169.77 (1C, C=O amide-pyridine) ppm. Anal. Calcd for C19H18ClN9O3 (455.86): C, 50.06; H, 3.98; Cl, 7.78; N, 27.65; O, 10.53%. Found C, 50.08; H, 4.00; Cl, 7.79; N, 27.63; O, 10.53%.

3.1.4. 7-Amino-6-(4-amino-5-mercapto-4H-1,2,4-triazol-3-yl)-5-(4-chlorophenyl)-1,3-dimethylpyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (4)

To form the mercapto triazole ring, dissolving of Compound 1 (3.88 g, 10 mmol) in 10 mL of NaOEt. Then, with continual stirring, hydrazine carbo thiohydrazide (1.06 g, 10 mmol) was added gradually. The mixture was then refluxed for 12 h at 80 °C. TLC confirmed that everything was going according to plan with the reaction. After cooling the precipitate to room temperature, it was collected, and cold methanol was used to thoroughly clean the solid. The crude product from dioxane was recrystallized to increase purity, and compound 4 was produced as a brownish-yellow powder.
Yield 76%; m.p. 296–298 °C; IR (νmax/cm−1): 3354, 3204 (2NH2), 3064 (CH-arom.), 2928 (CH-Aliph.), 2346 (SH), 1716, 1656 (2 C=O), 1597 (C=N pyridine), 1551, 1507 (2C=N pyrazole), 1491 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 3.06, 3.11 (s, 6H, 2CH3), 5.09 (s, 2H, NH2-pyrazole, D2O exchangeable), 7.25–7.57 (m, 4H, Ar-H), 7.93 (s, 2H, NH2-pyridine, D2O exchangeable), 13.45 (s, 1H, SH) ppm; 13C-NMR (DMSO-d6, 100 MHz); δ 23.09, 26.34 (2C, 2CH3), 102.24 (1C, C pyridine-C=O), 107.35 (1C, C pyridine-triazole), 127.54, 128.29, 128.95, 129.34, 130.42, 131.16 (6C, C-Ar), 152.04 (1C, C=N triazole-pyridine), 152.67 (1C, C pyridine-Ph-Cl), 154.81 (1C, C pyridine-N-CH3), 154.94, 158.84 (2C, 2C=O uracil), 155.82 (1C, C=N pyridine), 169.00 (1C, C=N triazole-SH) ppm; MS: m/z = 430.63 [M+], 186 (100%). Anal. Calcd for C17H15ClN8O2S (430.87): C, 47.39; H, 3.51; Cl, 8.23; N, 26.01; O, 7.43; S, 7.44%. Found C, 47.42; H, 3.59; Cl, 8.25; N, 26.00; O, 7.43; S, 7.44%.

3.1.5. 7-Amino-5-(4-chlorophenyl)-N-(hydrazinecarbonothioyl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine-6-carboxamide (5)

After dissolving Compound 1 (3.88 g, 10 mmol) in 10 mL of Gl. AcOH, the mixture was agitated at R.T. for 30 min. Thiosemicarbazide (0.92 g, 10 mmol) was added gradually while being continuously agitated once equilibrium was reached. The mixture was refluxed at 80 °C for 18 h, and progress was monitored by TLC (EtOAc/hexane, 1:1). After completion, the precipitate was filtered, washed with cold ethanol, and recrystallized from ethanol to give compound 5 as a dark yellow powder.
Yield 80%; m.p. ˃ 300 °C; IR (νmax/cm−1): 3441, 3281 (NH2-pyridine), 3349, 3254 (NH2-NH), 3201, 3127 (2NH), 3046 (CH- arom.), 2931 (CH-Aliph.), 1716, 1656 (2C=O uracil), 1597 (C=O amide-NH), 1560 (C=N), 1516, 1491 (C=C arom.), 1090 (C=S); 1H-NMR (DMSO-d6, 400 MHz,); δ 3.01 (s, 2H, NH2-NH, D2O exchangeable), 3.08, 3.10 (s, 6H, 2CH3), 7.30–7.91 (m, 4H, Ar-H), 7.93 (s, 2H, NH2-pyridine, D2O exchangeable), 10.89 (s, 1H, NH-NH2, D2O exchangeable), 12.06 (s, 1H, NH-CO, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz); δ 23.07, 26.37 (2C,2CH3), 105.24 (2C, 2C pyridine-C=O amide), 127.56, 129.36, 137.85 (6C, C-Ar), 151.40 (1C, C pyridine-Ph-Cl), 154.63, 158.94 (2C, C=O uracil), 157.28 (1C, C=N pyridine), 161.97 (1C, C pyridine-N-CH3), 171.62 (1C, C=O amide), 186.05 (1C, C=S) ppm. Anal. Calcd for C17H16ClN7O3S (433.87): C, 47.06; H, 3.72; Cl, 8.17; N, 22.60; O, 11.06; S, 7.39%. Found C, 47.16; H, 3.02; Cl, 8.17; N, 22.61; O, 11.04; S, 7.39%.

3.1.6. 7-Amino-5-(4-chlorophenyl)-6-(3,5-dimethyl-1H-pyrazole-1-carbonyl)-1,3-dimethylpyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (6)

Compound 2 (3.75 g, 10 mmol) and acetylacetone (1.02 mL, 10 mmol) were dissolved in 10 mL of sodium ethoxide solution and refluxed at 80 °C for 8 h. Reaction progress was checked by TLC. After cooling, the mixture was poured into 50 mL of ice-cold water, and the precipitate was collected by filtration, washed with cold methanol, and recrystallized from ethanol to yield compound 6 as orange crystals.
Yield 70%; m.p. 239–241 °C; IR (νmax/cm−1): 3384, 3304 (NH2), 3073 (CH- arom.), 2983, 2934, 2819, 2736 (CH-Alipha.), 1740 (C=O-pyrazole), 1688, 1650 (2C=O amide), 1591 (C=N pyridine), 1530 (C=N pyrazole), 1500 cm−1 (ν C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 2.71, 2.87 (s, 6H, 2CH3 pyrazole), 3.08, 3.12 (s, 6H, 2CH3 uracil), 6.24 (s, 1H, CH pyrazole), 7.31–8.41 (m, 4H, Ar-H), 8.49 (s, 2H, NH2, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz); δ 13.04, 13.97 (2C, 2CH3 pyrazole), 23.19, 26.67 (2C, 2CH3), 102.00, 104.11 (2C, 2C pyridine-C=O), 110.02 (1C, CH pyrazole), 127.27, 127.61, 128.02, 128.51, 129.49, 131.06 (6C, C-Ar), 141.64 (1C, C pyrazole-CH3), 151.06 (1C, C=N pyrazole), 154.93 (1C, C pyridine-Ph-Cl), 154.93, 158.83 (2C, C=O uracil), 161.79 (1C, C=N pyridine), 162.48 (1C, C pyridine-N-CH3), 165.81 (1C, C=O amide) ppm; MS: m/z = 438.87 [M+], 132 (100%). Anal. Calcd for C21H19ClN6O3 (430.87): C, 57.47; H, 4.36; Cl, 8.08; N, 19.15; O, 10.94%. Found C, 57.49; H, 4.41; Cl, 8.08; N, 19.13; O, 10.97%.

3.1.7. 7-Amino-5-(4-chlorophenyl)-6-(5-mercapto-1,3,4-oxadiazol-2-yl)-1,3-dimethylpyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (7)

By dissolving compound 2 (3.745 g, 10 mmol) in 8 mL of DMF, the mercapto-oxadiazole moiety formed. After adding carbon disulfide (0.6 g, 10 mmol), the reaction was conducted and refluxed with potassium hydroxide (KOH) acting as a catalyst for 4 h at 100 °C in a water bath. (TLC) was used to validate the reaction’s development. To induce precipitation, 50 mL of crushed ice was added to the reaction mixture after cooling to R.T. The resulting solid was recovered by vacuum filtering and treated with cold ethanol to eliminate residual impurities. Pure chemical 7 was created as a pale yellow crystalline powder by recrystallizing ethanol.
Yield (85%); m.p. 256–258 °C; IR (νmax/cm−1): 3411, 3347 (NH2), 3042 (CH-arom.), 2967, 2923 (CH-Aliph.), 2596 (SH), 1686, 1654 (2C=O amide), 1599, 1588 (2C=N oxadiazole), 1553 (C=N pyridine), 1465 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 3.07, 3.10 (s, 6H, 2CH3), 7.30–8.18 (m, 4H, Ar-H), 8.50 (s, 2H, NH2, D2O exchangeable), 13.80 (s, 1H, SH) ppm; 13C-NMR (DMSO-d6, 100 MHz); δ 23.00, 26.00 (2C, 2CH3), 102.94 (1C, C pyridine-C=O), 107.77 (1C, C pyridine-oxadiazole), 127.60, 128.71, 129.32, 129.64, 129.92, 131.37 (6C, C-Ar), 152.00 (1C, C pyridine-Ph-Cl), 154.43, 158.26 (2C, 2C=O uracil), 154.62 (1C, C=N pyridine), 162.22 (1C, C pyridine-N-CH3), 166.66 (1C, C=N oxadiazole-pyridine), 169.59 (1C, C=N triazole-SH) ppm. Anal. Calcd for C17H13ClN6O3S (416.84): C, 48.98; H, 3.14; Cl, 8.50; N, 20.16; O, 11.51; S, 7.69%. Found C, 49.01; H, 3.18; Cl, 8.52; N, 20.18; O, 11.51; S, 7.69%.

3.1.8. 3-Amino-5-[7-amino-5-(4-chlorophenyl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidin-6-yl]-4H-pyrazole-4-carbonitrile (8)

Compound 2 (3.75 g, 10 mmol) and malononitrile (0.58 mL, 10 mmol) were diluted in 10 mL of DMF and added. TLC was used to track the reaction’s progress over six hours of reflux at 90 °C. After finishing, the mixture was let to cool to room temperature, which caused a precipitate to form. Vacuum filtration was used to gather the solid product, and cold methanol was used to wash it. By recrystallizing an ethanol/water combination, pure chemical 8 was ultimately produced as an off-white powder.
Yield 77%; m.p. ˃ 300 °C; IR (νmax/cm−1): 3416, 3216 (2NH2), 3061 (CH-arom.), 2924, 2861 (CH-Alipha.), 2222 (CN), 1694, 1644 (2 C=O amide), 1597 (C=N pyridine), 1555, 1517 (2C=N pyrazole), 1504 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 2.79 (s, 1H, CH pyrazole), 3.08, 3.12 (s, 6H, 2CH3), 6.23 (s, 2H, NH2-pyrazole, D2O exchangeable), 7.29–8.40 (m, 4H, Ar-H), 8.49 (s, 2H, NH2-pyridine, D2O exchangeable) ppm;13C-NMR (DMSO-d6, 100 MHz); δ 23.06, 26.26 (2C, 2CH3), 33.50 (1C, CH pyrazole), 100.24 (1C, C pyridine-pyrazole), 103.39 (1C, C pyridine-C=O), 118.03 (1C, CN), 128.62,128.95, 129.45, 129.57, 130.48, 131.03 (6C, C-Ar), 152.69 (1C, C pyridine-Ph-Cl), 154.21, 158.00 (2C, C=O uracil), 157.17 (1C, C=N pyridine), 162.00 (1C, C pyridine-N-CH3), 165.80 (1C, C=N pyrazole-pyridine), 168.90 (1C, C=N pyrazole-NH2) ppm. Anal. Calcd for C19H15ClN8O2 (422.83): C, 53.97; H, 3.58; Cl, 8.38; N, 26.50; O, 7.57%. Found C, 53.99; H, 3.62; Cl, 8.41; N, 26.50; O, 7.59%.

3.1.9. 3-{5-[7-Amino-5-(4-chlorophenyl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido [2,3-d]pyrimidin-6-yl]-2-thioxo-1,3,4-oxadiazol-3(2H)-yl}propanenitrile (9)

Using triethylamine as a catalytic base, compound 7 (4.16 g, 10 mmol) and acrylonitrile (0.65 mL, 10 mmol) were mixed in 10 mL of DMF. (TLC) was used to track the mixture’s progress over a 6-hours reflux at 100 °C. Once finished, the reaction mixture was cooled to ambient temperature and subsequently added to 50 mL of ice-cold water, which caused a precipitate to form. To separate out any remaining contaminants, the crude product was collected via suction filtration and purified by washing with cold ethanol. Following recrystallization from DMF, pure compound 9 was produced as a brown powder.
Yield 75%; m.p. 229–231 °C; IR (νmax/cm−1): 3416, 3325 (NH2), 3028 (CH-arom.), 2989, 2945, 2920, 2873 (CH-Aliph.), 2249 (CN), 1686, 1666 (2C=O amide), 1619 (C=N oxadizole), 1599 (C=N pyridine), 1522 (C=C arom.), 1058 (C=S); 1H-NMR (DMSO-d6, 400 MHz): δ 3.09, 3.17 (s, 6H, 2CH3), 3.91 (t, J = 7.2 Hz, 2H, CH2-CN), 4.15 (t, J = 7.2 Hz, 2H, CH2), 7.28–7.63 (m, 4H, Ar–H), 8.51 (s, 2H, NH2, D2O exchangeable) ppm. 13C-NMR (DMSO-d6, 100 MHz); δ 16.10 (1C, CH2-CN), 23.02, 26.04 (2C, 2CH3), 45.70 (1C, CH2-oxadiazole), 100.04 (1C, C pyridine- oxadiazole), 102.24 (1C, C pyridine-C=O), 118.03 (1C, CN), 128.62, 128.8, 128.95, 129.57, 130.48, 131.03 (6C, C-Ar), 153.03 (1C, C pyridine-Ph-Cl), 154.90, 158.47 (2C, 2C=O uracil), 158.96 (1C, C=N pyridine), 159.26 (1C, C=N oxadiazole), 160.08 (1C, C pyridine-N-CH3), 174.00 (1C, C=S) ppm. Anal. Calcd for C20H16ClN7O3S (469.90): C, 51.12; H, 3.43; Cl, 7.54; N, 20.87; O, 10.21; S, 6.82%. Found C, 51.152; H, 3.48; Cl, 7.58; N, 20.89; O, 10.21; S, 6.82%.

3.1.10. 6-{4-[2-(2H-Tetrazol-5-yl)ethyl]-5-thioxo-4,5-dihydro-1,3,4-oxadiazol-2-yl}-7-amino-5-(4-chlorophenyl)-1,3-dimethylpyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (10)

To carry out the transformation (Tetrazole synthesis), compound 9 (4.16 g, 10 mmol) was stirred with sodium azide (0.65 g, 10 mmol) in 10 mL of DMF. A catalyst, ammonium chloride, was used to accelerate the reaction. TLC analysis was used to track the reaction’s progress during the mixture’s 10-hours reflux at 90 °C. Following completion, Once the reaction liquid had cooled to room temperature, 50 milliliters of crushed ice were added to cause precipitation. Compound 10 was obtained as a bright dark brown powder by vacuum filtering the solid, thoroughly cleaning it with cold ethanol to eliminate any leftover impurities, and then refining it via recrystallization from methanol.
Yield 66%; m.p. ˃ 300 °C; IR (νmax/cm−1): 3444, 3372 (NH2), 3282 (NH), 3072 (CH- arom.), 2961, 2832 (CH-Aliph.), 1683, 1636 (2C=O amide), 1603 (C=N oxadiazole), 1539 (C=N pyridine), 1491 (C=N tetrazole), 1466 (C=C arom.), 1297 (N=N), 1092 (C=S); 1H-NMR (DMSO-d6, 400 MHz), δ 3.07, 3.18 (s, 6H, 2CH3), 2.71, 4.16 (t, J = 7.0 Hz, 4H, 2CH2), 7.29–7.93 (m, 4H, Ar-H), 8.49 (s, 2H, NH2, D2O exchangeable), 14.00 (s, 1H, NH, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz) δ = 22.00 (1C, CH2-tetrazole), 23.06, 26.03 (2C, 2CH3), 54.89 (1C, CH2-N), 100.03 (1C, C pyridine-oxadiazole), 102.06 (1C, C pyridine-CO), 126.73, 127.84, 128.03, 129.78, 130.66, 131.00 (6C, C-Ar), 153.26 (1C, C pyridine-Ph-Cl), 154.46, 158.03 (2C, 2C=O uracil), 158.85 (1C, C=N pyridine), 159.44 (1C, C=N oxadiazole), 161.94 (1C, C uracil-NCH3), 165.94 (1C, C=N tetrazole), 174.12 (1C, C=S) ppm; MS: m/z = 512.93 [M+], 80 (100%). Anal. Calcd for C20H17ClN10O3S (512.93): C, 46.83; H, 3.34; Cl, 6.91; N, 27.31; O, 9.36; S, 6.25%. Found C, 46.83; H, 3.348; Cl, 6.961; N, 27.30; O, 9.36; S, 6.258%.

3.1.11. 7-Amino-5-(4-chlorophenyl)-1,3-dimethyl-6-(4-oxo-3a,5-dihydro-4H-pyrazolo [3,4-d]pyrimidin-3-yl)pyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (11)

Compound 8 (4.22 g, 10 mmol) and formic acid (0.4 mL, 10 mmol) was heated in a sand bath at 60 °C for 10 h. Following completion, the mixture was allowed to cool to ambient temperature before being added to 50 mL of crushed ice which caused the crude product to precipitate. The precipitate was gathered by vacuum filtering and meticulously washed with cold ethanol to remove impurities. After being purified from chloroform by recrystallization, compound 11 was obtained as a black powder.
Yield 68%; m.p. ˃ 300 °C; IR (νmax/cm−1): 3424, 3363 (NH2), 3161 (NH), 3060 (CH-arom.), 2974, 2923, 2870 (CH-Aliph.), 1723 (C=O lactam), 1683, 1666 (2C=O amide), 1635 (C=N pyrimidinone), 1585 (C=N pyridine), 1550, 1514 (2C=N pyrazole), 1483 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 2.32 (s, 1H, CH pyrazole), 3.06, 3.14 (s, 6H, 2CH3), 7.20–8.26 (m, 4H, Ar-H), 8.38 (s, 1H, CH pyrimidinone), 8.44 (s, 2H, NH2, D2O exchangeable), 11.44 (s, 1H, NH, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz); δ 23.06, 26.13 (2C, 2CH3), 46.83 (1C, CH pyrazole), 100.42 (1C, C pyridine-pyrazole), 103.09 (1C, C pyridine-C=O), 128.42, 128.58, 128.71, 129.44, 130.98, 134.13 (6C, C-Ar), 150.98 (1C, C=N pyrimidinone), 151.63 (1C, C pyridine-Ph-Cl), 154.82, 158.17 (2C, C=O uracil), 157.10 (1C, C=N pyridine), 162.81 (1C, C pyridine-N-CH3), 165.76 (1C, C=N pyrazole-pyridine), 169.85 (1C, C=O pyrimidinone), 181.01 (1C, C=N pyrazole- pyrimidinone) ppm; MS: m/z = 450 [M+], 319 (100%). Anal. Calcd for C20H15ClN8O3 (450.84): C, 53.28; H, 3.35; Cl, 7.86; N, 24.85; O, 10.65%. Found C, 53.30; H, 3.37; Cl, 7.89; N, 24.85; O, 10.66%.

3.1.12. 7-Amino-6-(4-amino-3aH-pyrazolo [3,4-d]pyrimidin-3-yl)-5-(4-chlorophenyl)-1,3-dimethylpyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (12)

Compound 8 (4.22 g, 10 mmol) and formamide (0.45 mL, 10 mmol) were fused for 12 h at 60 °C in a sand bath. After completing, the mixture was cooled to room temperature, which resulted in the appearance of a precipitate. The resulting solid was recovered by vacuum filtering and purifying by cold ethanol. A black powder representing pure component 12 was obtained when it was recrystallized from ethanol.
Yield 65%; m.p. 198–200 °C; IR (ν max/cm−1): 3422 (2NH2), 3067 (CH-arom.), 2964, 2919, 2821 (CH-Aliph.), 1702, 1646 (2C=O amide), 1610, 1594 (2C=N-dihydropyrimidin-4-amine), 1522 (C=N pyridine), 1517, 1492 (2C=N pyrazole), 1415 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 1.65 (s, 1H, CH pyrazole), 3.04, 3.12 (s, 6H, 2CH3), 7.14–8.14 (m, 4H, Ar-H), 8.25, 8.50 (s, 4H, 2NH2, D2O exchangeable), 8.91 (s, 1H, CH pyrimidine) ppm; 13C-NMR (DMSO-d6, 100 MHz) δ = 23.06, 26.23 (2C, 2CH3), 35.86 (1C, CH pyrazole), 100.10 (1C, C pyridine-pyrazole), 102.50 (1C, 1C pyridine-C=O), 127.22, 127.69, 128.47, 128.81, 130.99, 131.46 (6C, C-Ar), 152.07 (1C, C pyridine-Ph-Cl), 154.22, 158.42 (2C, C=O uracil), 156.62 (1C, C=N pyridine), 162.98 (1C, C pyridine-N-CH3), 162.42 (1C, C=N pyrimidine-NH2), 163.09 (1C, C=N pyrimidine), 167.12 (1C, C=N pyrazole-pyridine), 178.02 (1C, C=N pyrazole) ppm. Anal. Calcd for C20H16ClN9O2 (449.86): C, 53.40; H, 3.59; Cl, 7.88; N, 28.02; O, 7.11%. Found C, 53.42; H, 4.04; Cl, 7.90; N, 28.05; O, 7.11%.

3.1.13. 7-Amino-5-(4-chlorophenyl)-N-(3,5-dimethyl-1′H-[1,4′-bipyrazol]-3′-yl)-1,3-dimethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine-6-carboxamide (13)

To synthesize compound 13, a 50 mL rounded flask was filled with 10 mL of DMF, and compound 3 (4.55 g, 10 mmol) and acetylacetone (1.02 mL, 10 mmol) were dissolved in it. After adding 5 drops of triethylamine as a catalytic base, the mixture was refluxed for twelve hours at 90 °C. After finishing, the mixture was let to cool to R.T, which caused a precipitate to form. The separated solid was obtained by suction filtration and carefully washed with cold methanol to ensure purity. Pure compound 13 was obtained as an orange powder through recrystallization from ethanol.
Yield 69%; m.p. 268–270 °C; IR (νmax/cm−1): 3377, 3322 (NH2), 3275, 3250 (2NH), 3063 (CH-arom.), 2980, 2922, 2853 (CH-Aliph.), 1700, 1650 (2C=O uracil), 1596 (C=O amide), 1572, 1520, 1491 (3C=N), 1427 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,); δ 1.89, 2.10 (s, 6H, 2CH3 pyrazole), 3.02, 3.16 (s, 6H, 2CH3), 5.04 (s, 1H, CH methyl pyrazole), 7.24–7.46 (m, 4H, Ar-H), 7.83 (s, 1H, CH pyrazole), 8.41 (s, 2H, NH2, D2O exchangeable), 11.00 (s, 1H, NH-C=O, D2O exchangeable), 12.22 (s, 1H, NH pyrazole, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz); δ 12.32, 13.50 (2C, 2CH3 pyrazole), 23.16, 26.29 (2C, 2CH3), 90.00 (1C, C pyrazole-di methyl pyrazole), 102.22 (2C, 2C pyridine-C=O), 112.25 (1C, CH di methyl pyrazole), 127.52, 127.72, 128.4, 129.58, 130.41, 131.76 (6C, C-Ar), 138.60 (1C, CH pyrazole), 142.20 (1C, C=N methyl pyrazole), 155.35 (1C, 1C di methyl pyrazole), 154.38, 158.26 (2C, 2C=O uracil), 156.03 (1C, C=N pyrazole), 156.76 (2C, 2C=C fused), 162.26 (1C, C=N pyridine), 169.96 (1C, C=O amide) ppm; MS: m/z = 519 [M+], 300 (100%). Anal. Calcd for C24H22ClN9O3 (519.95): C, 55.44; H, 4.27; Cl, 6.82; N, 24.25; O, 9.23%. Found C, 55.42; H, 4.27; Cl, 6.84; N, 24.25; O, 9.25%.

3.1.14. 7-Amino-6-(6-chloro-7H-[1,2,4]triazolo [3,4-b][1,3,4]thiadiazin-3-yl)-5-(4-chlorophenyl)-1,3-dimethylpyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (14)

In a 100 mL round-bottom flask, compound 4 (4.31 g, 10 mmol) was agitated with 0.8 mL of chloroacetyl chloride (10 mmol) in 10 mL of ethanol, while catalyst such as tri ethyl amine (five drops) was added. At room temperature, the reaction mixture was agitated for twenty-four hours. After finishing, the solid product was recovered through vacuum filtration, and cold ethanol was used for washing to remove residual impurities. Recrystallization from diethyl ether afforded pure compound 14 as a dark red powder.
Yield 88%; m.p. 298–300 °C; IR (νmax/cm−1): 3410, 3312 (NH2), 3050 (CH-arom.), 2929 (CH-Aliph.), 1691, 1648 (2C=O amide), 1595 (C=N pyridine), 1558, 1512 (2C=N triazole), 1491 (C=N thiadiazin), 1453 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz); δ 3.06, 3.15 (s, 6H, 2CH3), 4.92 (s, 2H, CH2), 7.26–7.41 (m, 4H, Ar-H), 8.60 (s, 2H, NH2, D2O exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz): δ 23.07, 26.34 (2C, 2CH3), 34.16 (1C, CH2 thiadiazine), 102.80 (1C, C pyridine-C=O), 106.00 (1C, C pyridine-triazole), 128.65, 128.85, 128.92, 129.31, 129.83, 130.84 (6C, C-Ar), 150.36 (C=N triazole-pyridine), 151.06 (1C, C=N thiadiazine), 151.08 (1C, C pyridine-Ph-Cl), 154.88, 159.36 (2C, 2C=O uracil), 157.22 (1C, C pyridine-N-CH3), 157.50 (1C, C=N pyridine), 162.80 (1C, C=N triazole- thiadiazine) ppm. Anal. Calcd for C19H14Cl2N8O2S (489.34): C, 46.64; H, 2.88; Cl, 14.49; N, 22.90; O, 6.54; S, 6.55%. Found C, 46.65; H, 3.00; Cl, 14.52; N, 22.91; O, 6.53; S, 6.55%.

3.1.15. 7-Amino-5-(4-chlorophenyl)-6-(6-mercapto-[1,2,4]triazolo[3,4-b][1,3,4]thiadiazol-3-yl)-1,3-dimethylpyrido[2,3-d]pyrimidine-2,4(1H,3H)-dione (15)

Compound 4 (4.30 g, 10 mmol) and carbon disulfide (0.6 g, 10 mmol) were combined in 10 mL of pyridine to create compound 15. The reaction was refluxed for 4 h at 60 °C. After finishing, neutralization with HCl and 50 mL of ice-cold water was performed with cooled mixture. Cold ethanol was used to remove impurities from the precipitate after it had been isolated by vacuum filtering. Compound 15 was refined by recrystallization from ethanol, and the final result was a purple powder.
Yield 65%; m.p. 239–241 °C; IR (νmax/cm−1): 3449, 3276 (NH2), 3072, 3049 (CH arom.), 2984, 2940 (CH-Aliph.), 2550 (SH), 1683, 1633 (2 C=O amide), 1584, 1524 (4 C=N), 1432 (C=C arom.); 1H-NMR (DMSO-d6, 400 MHz,): δ 3.07, 3.13 (s, 6H, 2CH3), 7.15–7.70 (m, 4H, Ar-H), 8.52 (s, 2H, NH2, D2O exchangeable), 13.24 (s, 1H, SH) ppm; 13C-NMR (DMSO-d6, 100 MHz) δ = 23.04, 26.20 (2C, 2CH3), 102.00 (1C, C pyridine-C=O), 107.57 (1C, C pyridine-triazole), 128.68, 129.07, 129.74, 131, 131.26, 132.28 (6 C, C-Ar), 152.00 (1C, C=N triazole-pyridine), 152.89 (1C, C pyridine-Ph-Cl), 154.13 (1C, C uracil-N), 154.93, 158.26 (2C, 2C=O uracil), 155.58 (1C, C pyridine-NH2), 166.07 (1C, C=N triazol-S), 188.17 (1C, C=N-SH) ppm; MS: m/z = 572 [M+], 284 (100%). Anal. Calcd for C18H13ClN8O2S2 (472.93): C, 45.71; H, 2.77; Cl, 7.50; N, 23.69; O, 6.77; S, 13.56%. Found C, 47.71; H, 2.77; Cl, 7.54; N, 23.69; O, 6.78; S, 13.56%.

3.1.16. 7-Amino-5-(4-chlorophenyl)-1,3-dimethyl-N-(3-methyl-5-oxo-4,5-dihydro-1H-pyrazole-1-carbonothioyl)-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine-6-carboxamide (16)

A new methyl pyrazol-3-one derivative 16, was synthesized by refluxing of compound 6 (4.33 g, 10 mmol) with ethyl acetoacetate (1.3 mL, 10 mmol) in 10 mL of DMF, activation was made by sodium hydride (NaH, 0.24 g, 10 mmol) as a base for 14 h at 100 °C, with thin-layer chromatography (TLC) tracking the development. Following completion, introducing precipitate after cooling was made by adding 50 milliliters of crushed ice. Following vacuum filtration, the solid was collected, and any impurities were removed by washing it with cold ethanol. By recrystallizing ethanol, pure compound 16 was produced as a pale orange powder.
Yield 79%; m p 229–231 °C; IR (νmax/cm−1): 3341, 3194 (NH2), 3061 (CH-arom.), 2931, 2839, 2816 (CH-Aliph.), 1727 (C=O pyrazalone), 1703, 1652 (2C=O uracil), 1597 (C=O amide), 1560 (C=N pyridine), 1516 (C=N pyrazalone), 1491 (C=C arom.), 1090 (C=S). 1H-NMR (DMSO-d6, 400 MHz,) δ = 1.94 (s, 3H, CH3), 3.01 (s, 2H, CH2), 3.06, 3.16 (s, 6H, 2CH3), 7.14–7.39 (m, 4H, Ar-H), 8.30 (brs, 2H, NH2 exchangeable), 12.42 (brs, 1H, NH exchangeable) ppm; 13C-NMR (DMSO-d6, 100 MHz) δ = 16.00 (1C, CH3 pyrazalone), 23.32, 26.13 (2C, 2CH3), 45.05 (1C, CH2), 103.60 (2C, 2C pyridine-C=O), 127.52, 128.07, 129.07, 129.33, 129.84, 131.64 (6C, C-Ar), 151.34 (1C, C pyridine-Ph-Cl), 154.58, 158.89 (2C, C=O uracil), 159.58 (2C, 2C=N pyridine, pyrazole), 161.25 (1C, C pyridine-N-CH3), 165.33 (1C, 1C=O pyrazole), 170.06 (1C, C=O amide), 182.25 (1C, C=S) ppm; MS: m/z = 499 [M+], 245 (100%). Anal. Calcd for C21H18ClN7O4S (499.93): C, 50.45; H, 3.63; Cl, 7.09; N, 19.61; O, 12.80; S, 6.41%. Found C, 50.48; H, 3.70; Cl, 7.12; N, 19.61; O, 12.82; S, 6.41%.

3.2. Biological Activity

The supplementary material included all descriptions of all in vitro investigations.

3.2.1. Assessment of Antiproliferative Activity

Three human malignant cells were acting as evaluating agents of the antitumor efficacy of the newly obtained 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine -based derivatives 116 in vitro in order to determine their IC50 values.

3.2.2. In Vitro Enzyme Inhibitory Assay Against EGFRWT, Mutant EGFRL858R, and EGFRT790M

The inhibitory effect of newly 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives 1 and 2 against EGFRWT, mutant EGFRL858R, and EGFRT790M was investigated in vitro in comparison to erlotinib.

3.2.3. Cell Cycle and Apoptosis Analysis

Apoptosis research and cell cycle analysis were used on malignant MCF-7 cells for 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivative 1 through flow cytometry.

3.2.4. In Vitro Wound Healing Assessment

The screening procedure and details made use of the wound healing (scratch assay) were provided in the ESI materials.

3.2.5. In Vitro Antibacterial Assessment

Using the agar plate diffusion technique, the inhibition zone widths of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine -based derivatives 116 were examined against a number of species. The two-fold serial dilution method was used to further assess the minimum inhibitory concentration (MIC).

3.3. Computational Studies

3.3.1. ADMET In Silico Evaluation

The admetSAR 1.0 free web tools can be used to analyze the ADMET evaluation of the targeted drugs and SwissADME.

3.3.2. Docking Simulation

ChemDraw version 12 was used to draw Compounds 1 and 2, which were then transformed to 3D and energy-minimized in MOE (v.2024.0601). The Protonate 3D procedure was used to build the EGFRWT and EGFRT790M crystal structures (PDB: 1M17, 3IKA) for docking upon collection from the Protein Data Bank.
Prior to evaluating the compounds in the ATP-binding sites, Triangle Matcher placement and the London dG score function were employed for docking, which was confirmed by re-docking native ligands. The same procedure was used to produce and dock HCV NS5A (PDB: 1AJ6, 3FQM) and E. coli DNA gyrase. The Supplementary Information contains all of the experimental details.

3.3.3. Molecular Dynamics Simulation

The supplementary file contains the specifics of the molecular dynamics simulation of 1,3-dimethyluracil derivative 1 with EGFRWT and mutant EGFRT790M.

3.3.4. Quantum Chemical Calculation

Understanding molecular characteristics and reaction pathways requires the use of quantum mechanical computations. In this study, density functional theory (DFT) calculations were performed using the 6311G++(d,p) basis set and the B3LYP functional to optimize molecular structures and analyze electronic properties. Gaussian software was used for all computations; more information is given in the associated documentation.

4. Conclusions

The outcomes of this study highlight the 7-substituted tetrahydropyrido[2,3-d]pyrimidine-6-carboxylates as a versatile chemotype with significant anticancer and antibacterial properties, accompanied by computationally derived antiviral activity predictions. Compound 1, in particular, demonstrated multi-target activity through EGFR inhibition (wild and mutant forms), apoptosis induction, and antibacterial efficacy. These findings encourage further structural optimization to improve potency and selectivity, alongside evaluation in resistant cancer models and advanced microbial strains. In addition, computational ADMET predictions provided insights into pharmacokinetic and toxicity profiles. The antiviral potential, assessed exclusively via computational approaches, highlights possible avenues for further experimental validation. Collectively, these results support the therapeutic promise of these derivatives in oncology and infectious disease research.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph18101472/s1, The following supporting information contain IC50 values of compounds against HeLa, HepG-2, and MCF-7 cell lines, Tables S1–S9, and Figures S1–S56.

Author Contributions

H.S.E.-H.: conceptualization, supervision, data interpretation, manuscript writing, and final review. H.E.S.: experimental work, funding acquisition (chemicals), and manuscript review. M.A.H.: supervision, conceptualization, manuscript writing, and review. E.S.N.: anticancer activity experiments, docking studies, anticancer section writing, and review. A.F.E.-S.: antimicrobial activity experiments, results writing, docking studies, manuscript writing, and review. N.A.A. and Asmaa Saleh, A.S. (Amr Sbry): project funding and manuscript review. M.F.H.: quantum chemical calculations, related section writing, and review. A.S. (Asmaa Saleh): experimental work, results writing, and manuscript review. A.A.-H.A.-R.: supervision, conceptualization, and manuscript review. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2025R89), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors gratefully acknowledge the financial support from Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2025R89), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia. The authors also extend their sincere thanks to the Faculty of Science, Menoufia University, Egypt, where the experimental work was conducted.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bray, F.; Laversanne, M.; Sung, H.; Ferlay, J.; Siegel, R.L.; Soerjomataram, I.; Jemal, A. Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2024, 74, 229–263. [Google Scholar] [CrossRef]
  2. Filho, A.M.; Laversanne, M.; Ferlay, J.; Colombet, M.; Piñeros, M.; Znaor, A.; Parkin, D.M.; Soerjomataram, I.; Bray, F. The GLOBOCAN 2022 cancer estimates: Data sources, methods, and a snapshot of the cancer burden worldwide. Int. J. Cancer 2025, 156, 1336–1346. [Google Scholar] [CrossRef]
  3. Bizuayehu, H.M.; Ahmed, K.Y.; Kibret, G.D.; Dadi, A.F.; Belachew, S.A.; Bagade, T.; Tegegne, T.K.; Venchiarutti, R.L.; Kibret, K.T.; Hailegebireal, A.H.; et al. Global disparities of cancer and its projected burden in 2050. JAMA Netw. Open 2024, 7, e2443198. [Google Scholar] [CrossRef]
  4. Ge, M.; Chen, X.Y.; Huang, P.; Fleishman, J.S.; Yang, D.H.; Wu, Z.X.; Chen, Z.S. Understanding and overcoming multidrug resistance in cancer. Nat. Rev. Clin. Oncol. 2025, 22, 760–780. [Google Scholar] [CrossRef]
  5. Yao, L.; Jiang, B.; Xu, D. Strategies to combat cancer drug resistance: Focus on copper metabolism and cuproptosis. Cancer Drug Resist. 2025, 8, 15. [Google Scholar] [CrossRef]
  6. Wu, G.Y.; Xiao, M.Z.; Hao, W.C.; Yang, Z.S.; Liu, X.R.; Xu, D.S.; Peng, Z.X.; Zhang, L.Y. Drug resistance in breast cancer: Mechanisms and strategies for management. Drug Resist. Updat. 2025, 83, 101288. [Google Scholar] [CrossRef]
  7. Shyamsunder, S.; Lu, Z.; Takiar, V.; Waltz, S.E. Challenges and resistance mechanisms to EGFR targeted therapies in head and neck cancers and breast cancer: Insights into RTK dependent and independent mechanisms. Oncotarget 2025, 16, 508. [Google Scholar] [CrossRef]
  8. Bouchard, N.; Daaboul, N. Lung cancer: Targeted therapy in 2025. Curr. Oncol. 2025, 32, 146. [Google Scholar] [CrossRef]
  9. da Silva Santos, E.; Nogueira, K.A.B.; Fernandes, L.C.C.; Martins, J.R.P.; Reis, A.V.F.; Neto, J.D.B.V.; da Silva Júnior, I.J.; Pessoa, C.; Petrilli, R.; Eloy, J.O. EGFR targeting for cancer therapy: Pharmacology and immunoconjugates with drugs and nanoparticles. Int. J. Pharm. 2021, 592, 120082. [Google Scholar] [CrossRef]
  10. Khan, M.N.; Farooq, U.; Khushal, A.; Wani, T.A.; Zargar, S.; Khan, S. Unraveling potential EGFR kinase inhibitors: Computational screening, molecular dynamics insights, and MMPBSA analysis for targeted cancer therapy development. PLoS ONE 2025, 20, e0321500. [Google Scholar] [CrossRef]
  11. Pan, Q.; Lu, Y.; Xie, L.; Wu, D.; Liu, R.; Gao, W.; Luo, K.; He, B.; Pu, Y. Recent advances in boosting EGFR tyrosine kinase inhibitors-based cancer therapy. Mol. Pharm. 2023, 20, 829–852. [Google Scholar] [CrossRef]
  12. Shen, Q.; Tang, H.; Chen, S. EGFR-TKIs for lung cancer treatment: Development, application, and side effects. Front. Oncol. 2025, 15, 1617788. [Google Scholar] [CrossRef]
  13. Soliman, S.M.; Abdel-Rahman, A.A.; Nossier, E.S.; Hussein, M.F.; Sabry, A.; El-Hema, H.S. Design, antiproliferative potency, and in silico studies of novel 5-(methylfuran-3-yl)thio)-3-phenylquinazolin-4(3H)-one based derivatives as potential EGFR inhibitors. Sci. Rep. 2025, 15, 27992. [Google Scholar] [CrossRef]
  14. Popper, H.H.; Ryska, A.; Timar, J.; Olszewski, W. Molecular pathology of lung cancer: Current status and perspectives. Transl. Lung Cancer Res. 2014, 3, 291–300. [Google Scholar]
  15. Thomas, R.; Weihua, Z. Rethinking EGFR biology: Impact on anti-EGFR cancer therapeutics. Front. Oncol. 2019, 9, 800. [Google Scholar]
  16. Sigismund, S.; Avanzato, D.; Lanzetti, L. Emerging functions of the EGFR in cancer. Mol. Oncol. 2018, 12, 3–20. [Google Scholar] [CrossRef]
  17. Su, C.; Sun, S.Y. Fourth-generation epidermal growth factor receptor-tyrosine kinases inhibitors: Hope and challenges. Transl. Cancer Res. 2024. 13, 3929–3934. [CrossRef]
  18. Campos, J.F.; Besson, T.; Berteina-Raboin, S. Review on the Synthesis and Therapeutic Potential of Pyrido[2,3-d], [3,2-d], [3,4-d] and [4,3-d]pyrimidine Derivatives. Pharmaceuticals 2022, 15, 352. [Google Scholar] [CrossRef]
  19. El Malah, T.; Abdel Mageid, R.E.; Shamroukh, A.H. Review of the Synthesis, Reactions and Pharmaceutical Potential of Pyrido[2,3-d]Pyrimidine Derivatives. Curr. Org. Chem. 2023, 27, 260–281. [Google Scholar] [CrossRef]
  20. Bandi, S.R.; Kapavarapu, R.; Palabindela, R.; Azam, M.; Min, K.; Narsimha, S. Synthesis of novel pyrido [2, 3-d] pyrimidine-thiazolidine-1, 2, 3-triazoles: Potent EGFR targeting anticancer agents. J. Mol. Struct. 2023, 1294, 136451. [Google Scholar] [CrossRef]
  21. Fu, J.; Yu, J.; Zhang, X.; Chang, Y.; Fan, H.; Dong, M.; Li, M.; Liu, Y.; Hu, J. Design, synthesis, and biological evaluation of pyrido [2, 3-d] pyrimidine and thieno [2, 3-d] pyrimidine derivatives as novel EGFRL858R/T790M inhibitors. J. Enzyme Inhib. Med. Chem. 2023, 38, 2205605. [Google Scholar] [CrossRef]
  22. Kumar, A.; Backer, N.; Paliwal, H.; Singh, A.K.; Debbaraman, T.; Singh, V.; Kumar, P. Synthesis and anticancer evaluation of diaryl pyrido [2, 3-d] pyrimidine/alkyl substituted pyrido [2, 3-d] pyrimidine derivatives as thymidylate synthase inhibitors. BMC 2024, 18, 161. [Google Scholar] [CrossRef]
  23. García-Marín, J.; Griera, M.; Sánchez-Alonso, P.; Di Geronimo, B.; Mendicuti, F.; Rodríguez-Puyol, M.; Alajarín, R.; de Pascual-Teresa, B.; Vaquero, J.J.; Rodríguez-Puyol, D. Pyrrolo[1,2-a]quinoxalines: Insulin mimetics that exhibit potent and selective inhibition against protein tyrosine phosphatase 1B. ChemMedChem 2020, 15, 1788–1801. [Google Scholar] [CrossRef]
  24. Shaha, C.; Bag, K.; Gangjee, A. A Review on Fused Pyrimidine Systems as EGFR Inhibitors and Their Structure–Activity Relationship. Front. Chem. 2022, 10, 861288. [Google Scholar] [CrossRef]
  25. Kumar, A.; Singh, P.; Singh, H. Synthesis and anticancer evaluation of novel pyrrole-pyrido[2,3-d]pyrimidine-based compounds as thymidylate synthase inhibitors. Chem. Biol. Drug Des. 2023, 101, 456–468. [Google Scholar] [CrossRef]
  26. Edupuganti, R.; Wang, Q.; Tavares, C.D.; Chitjian, C.A.; Bachman, J.L.; Ren, P.; Anslyn, E.V.; Dalby, K.N. Synthesis and biological evaluation of pyrido [2, 3-d] pyrimidine-2, 4-dione derivatives as eEF-2K inhibitors. Bioorg. Med. Chem. 2014, 22, 4910–4916. [Google Scholar] [CrossRef]
  27. Gao, X.; Cen, L.; Li, F.; Wen, R.; Yan, H.; Yao, H.; Zhu, S. Biochemical and biophysical evaluation of pyrido[2,3-d]pyrimidine derivatives. Biochem. Biophys. Res. Commun. 2018, 505, 761–767. [Google Scholar] [CrossRef]
  28. Jahanshahi, P.; Mamaghani, M.; Haghbin, F.; Hossein Nia, R.; Rassa, M. Synthesis and structural analysis of novel pyrido[2,3-d]pyrimidine derivatives. J. Mol. Struct. 2018, 1155, 520–529. [Google Scholar] [CrossRef]
  29. Naresh Kumar, R.; Dev, G.J.; Ravikumar, N.; Krishna Swaroop, D.; Debanjan, B.; Bharath, G.; Narsaiah, B.; Jain, S.N.; Gangagni Rao, A. Synthesis and anticancer evaluation of novel pyridopyrimidines. Bioorg. Med. Chem. Lett. 2016, 26, 2927–2930. [Google Scholar] [CrossRef] [PubMed]
  30. Hurlbert, B.; Valenti, B. 1,4-Dihydropyridines. II. 3,5-Dicarboxylic acid derivatives with hypotensive activity. J. Med. Chem. 1968, 11, 708–710. [Google Scholar] [CrossRef]
  31. Heber, D.; Heers, C.; Ravens, U. Cardiac effects of novel pyridopyrimidines with calcium antagonist activity. Pharmazie 1993, 48, 537–541. [Google Scholar]
  32. Keshk, R.M.; Salama, Z.A.; Elsaedany, S.K.; ElRehim, E.M.A.; Beltagy, D.M. Synthesis, antimicrobial, anti-inflammatory, antioxidant and cytotoxicity of new pyrimidine and pyrimidopyrimidine derivatives. Sci. Rep. 2025, 15, 9328. [Google Scholar] [CrossRef]
  33. Youssef, N.A.; Selim, Y.A.; Shaheen, L.A.; Fadda, A.A.; Gaffer, H.E.; El-Hadidy, S.A. Design and synthesis of novel pyridopyrimidine moieties linked to coumarin ring of potentially active cytotoxic agents and their docking study. Sci. Rep. 2025, 15, 20155. [Google Scholar] [CrossRef]
  34. Abu-Hashem, A.A.; Al-Hussain, S.A. Design, synthesis, antimicrobial activity, and molecular docking of novel thiazoles, pyrazoles, 1,3-thiazepinones, and 1,2,4-triazolopyrimidines derived from quinoline-pyrido[2,3-d]pyrimidinones. Pharmaceuticals 2024, 17, 1632. [Google Scholar] [CrossRef]
  35. Hammouda, M.M.; Rashed, M.M.; Elattar, K.M.; Osman, A.M. Synthetic strategies of heterocycle-integrated pyridopyrimidine scaffolds supported by nano-catalysts. RSC Adv. 2023, 13, 11600–11634. [Google Scholar] [CrossRef]
  36. Abdelgawad, M.A.; Al-Sanea, M.M.; Musa, A.; Elmowafy, M.; El-Damasy, A.K.; Azouz, A.A.; Bakr, R.B. Docking study, synthesis, and anti-inflammatory potential of some new pyridopyrimidine-derived compounds. J. Inflamm. Res. 2022, 15, 15–451. [Google Scholar] [CrossRef] [PubMed]
  37. Feng, Y.; Liu, G.; Zhang, C.; Li, J.; Li, Y.; Liu, L. Fluorescent immunoassay with a copper polymer as the signal label for catalytic oxidation of O-phenylenediamine. Molecules 2022, 27, 3675. [Google Scholar] [CrossRef] [PubMed]
  38. DeGoey, D.A.; Betebenner, D.A.; Grampovnik, D.J.; Liu, D.; Pratt, J.K.; Tufano, M.D.; He, W.; Krishnan, P.; Pilot-Matias, T.J.; Marsh, K.C.; et al. Discovery of novel NS5A inhibitors for hepatitis C virus. Bioorg. Med. Chem. Lett. 2013, 23, 3627–3630. [Google Scholar] [CrossRef] [PubMed]
  39. Tandel, H.T.; Patel, S.K. Microwave assisted synthesis and antimicrobial activity of pyridopyrimidines. Indian J. Chem. B 2020, 59, 502–508. [Google Scholar]
  40. Zohny, Y.M.; Awad, S.M.; Rabie, M.A.; Alsaidan, O.A. Design, synthesis, molecular modeling, and biological evaluation of novel pyrimidine derivatives as potential calcium channel blockers. Molecules 2023, 28, 4869. [Google Scholar] [CrossRef]
  41. Mohire, P.P.; Chandam, D.R.; Patravale, A.A.; Choudhari, P.; Karande, V.; Ghosh, J.S.; Deshmukh, M.B. Design and synthesis of new pyridopyrimidine-based compounds with anticancer potential. Polycycl. Aromat. Compd. 2020, 42, 137–155. [Google Scholar] [CrossRef]
  42. D’agostino, L.A.; Sjin, R.T.T.; Niu, D.; McDonald, J.J.; Zhu, Z.; Liu, H.; Mazdiyasni, H.; Petter, R.C.; Singh, J.; Barrague, M.; et al. Substituted Pyrido[2,3-d]Pyrimidines as Inhibitors of Protein Kinases. US Patent 10065966, 4 September 2018. [Google Scholar]
  43. Kots, A.Y.; Choi, B.K.; Estrella-Jimenez, M.E.; Warren, C.A.; Gilbertson, S.R.; Guerrant, R.L.; Murad, F. Pyridopyrimidine derivatives as inhibitors of cyclic nucleotide synthesis: Application for treatment of diarrhea. Proc. Natl. Acad. Sci. USA 2008, 105, 8440–8445. [Google Scholar] [CrossRef]
  44. Dasari, S.R.; Tondepu, S.; Vadali, L.R.; Seelam, N. Synthesis and characterization of novel pyridopyrimidine derivatives. Synth. Commun. 2020, 50, 2950–2961. [Google Scholar] [CrossRef]
  45. Yadav, P.; Shah, K. An overview on synthetic and pharmaceutical prospective of pyrido[2,3-d]pyrimidines scaffold. Chem. Biol. Drug Des. 2020, 97, 633–648. [Google Scholar] [CrossRef]
  46. Wei, X.; Fan, D.; Dong, H.; Li, M.; Lv, Q.; Zheng, P.; Wang, L. Advances in the synthesis and SAR of pyrido[2,3-d]pyrimidine scaffold. Curr. Pharm. Des. 2025, 31, 1521–1536. [Google Scholar] [CrossRef] [PubMed]
  47. Jubete, G.; Puig de la Bellacasa, R.; Estrada-Tejedor, R.; Teixidó, J.; Borrell, J.I. Pyrido[2,3-d]pyrimidin-7(8H)-ones: Synthesis and biomedical applications. Molecules 2019, 24, 4161. [Google Scholar] [CrossRef] [PubMed]
  48. Elsebaie, H.A.; El-Moselhy, T.F.; El-Bastawissy, E.A.; Elberembally, K.M.; Badi, R.M.; Elkaeed, E.B.; Shaldam, M.A.; Eldehna, W.M.; Tawfik, H.O. Development of new thieno[2,3-d]pyrimidines as dual EGFR and STAT3 inhibitors endowed with anticancer and pro-apoptotic activities. Bioorg. Chem. 2024, 143, 107101. [Google Scholar] [CrossRef] [PubMed]
  49. Kurumurthy, C.; Sambasiva, P.; Veeraswamy, B.; Santhoshkumar, G.; Shanthan, P.; Narsaiah, B.; Velatooru, L.R.; Pamanji, R.; Venkateswara, J. Synthesis of novel alkyltriazole tagged pyrido[2,3-d]pyrimidine derivatives and their anticancer activity. Eur. J. Med. Chem. 2011, 46, 3462–3470. [Google Scholar] [CrossRef]
  50. Dongre, R.S.; Meshram, J.S.; Selokar, R.S.; Almalki, F.A.; Hadda, T.B. Antibacterial activity of synthetic pyrido[2,3-d]pyrimidines armed with nitrile groups: POM analysis and identification of pharmacophore sites of nitriles as important pro-drugs. New J. Chem. 2018, 42, 15610–15617. [Google Scholar] [CrossRef]
  51. Zhang, H.J.; Wang, S.B.; Wen, X.; Li, J.Z.; Quan, Z.S. Design, synthesis, and evaluation of the anticonvulsant and antidepressant activities of pyrido[2,3-d]pyrimidine derivatives. Med. Chem. Res. 2016, 25, 1287–1298. [Google Scholar] [CrossRef]
  52. Nemr, M.T.; AboulMagd, A.M. New fused pyrimidine derivatives with anticancer activity: Synthesis, topoisomerase II inhibition, apoptotic inducing activity and molecular modeling study. Bioorg. Chem. 2020, 103, 104134. [Google Scholar] [CrossRef]
  53. Sayed, M.T.E.L.; Hussein, H.A.R.; Elebiary, N.M.; Hassan, G.S.; Elmessery, S.M.; Elsheakh, A.R.; Nayel, M.; Abdel-Aziz, H.A. Tyrosine kinase inhibition effects of novel pyrazolo [1,5-a]pyrimidines and pyrido[2,3-d]pyrimidines ligand: Synthesis, biological screening and molecular modeling studies. Bioorg. Chem. 2018, 78, 312–323. [Google Scholar] [CrossRef]
  54. Eissa, A.A.; Aljamal, K.F.; Ibrahim, H.S.; Allam, H.A. Design and synthesis of novel pyridopyrimidine derivatives with anchoring non-coplanar aromatic extensions of EGFR inhibitory activity. Bioorg. Chem. 2021, 116, 105318. [Google Scholar] [CrossRef]
  55. Kumar, A.; Bhagat, K.K.; Singh, A.K.; Singh, H.; Angre, T.; Verma, A.; Kumar, P. Medicinal chemistry perspective of pyrido[2,3-d]pyrimidines as anticancer agents. RSC Adv. 2023, 13, 6872–6908. [Google Scholar] [CrossRef]
  56. Yu, L.; Huang, M.; Xu, T.; Tong, L.; Yan, X.E.; Zhang, Z.; Xu, Y.; Yun, C.; Xie, H.; Ding, K.; et al. A structure-guided optimization of pyrido[2,3-d]pyrimidin-7-ones as selective inhibitors of EGFR L858R/T790M mutant with improved pharmacokinetic properties. Eur. J. Med. Chem. 2017, 126, 1107–1117. [Google Scholar] [CrossRef] [PubMed]
  57. Gaber, A.A.; Bayoumi, A.H.; El-Morsy, A.M.; Sherbiny, F.F.; Mehany, A.B.; Eissa, I.H. Design, synthesis and anticancer evaluation of 1H-pyrazolo[3,4-d]pyrimidine derivatives as potent EGFR WT and EGFR T790M inhibitors and apoptosis inducers. Bioorg. Chem. 2018, 80, 375–395. [Google Scholar] [CrossRef] [PubMed]
  58. Lin, Z.; Wang, Q.; Jiang, T.; Wang, W.; Zhao, J.J. EGFR mutations in non-small cell lung cancer. Oncol. Lett. 2025, 30, 375. [Google Scholar] [CrossRef]
  59. Elzahabi, H.S.A.; Nossier, E.S.; Alasfoury, R.A.; El-Manawaty, M.; Sayed, S.M.; Elkaeed, E.B.; Metwaly, A.M.; Hagras, M.; Eissa, I.H. Design, synthesis, and anti-cancer evaluation of new pyrido[2,3-d]pyrimidin-4(3H)-one derivatives as potential EGFRWT and EGFRT790M inhibitors and apoptosis inducers. J. Enzym. Inhib. Med. Chem. 2022, 37, 1053–1076. [Google Scholar] [CrossRef]
  60. Yang, F.; Liu, J.; Xu, M.; Peng, B. Acquired multiple EGFR mutations-mediated resistance to a third-generation tyrosine kinase inhibitor in a patient with lung adenocarcinoma who responded to afatinib: A case report and literature review. Transl. Lung Cancer Res. 2024, 13, 617–627. [Google Scholar] [CrossRef]
  61. Wu, W.; Yu, S.; Huang, J.; Qi, Q.; Wu, Y.; Dong, J. Molecular heterogeneity and treatment outcome of EGFR exon 20 insertion mutations in Chinese patients with advanced non-small cell lung cancer: Insights from a large-scale real-world study. BMC Cancer 2024, 24, 1010. [Google Scholar] [CrossRef]
  62. Satyanarayana, G.; Enriquez, K.T.; Sun, T.; Klein, E.J.; Abidi, M.; Advani, S.M.; Awosika, J.; Bakouny, Z.; Bashir, B.; Berg, S.; et al. Coinfections in Patients with Cancer and COVID-19: A COVID-19 and Cancer Consortium (CCC19) Study. Open Forum Infect. Dis. 2022, 9, ofac037. [Google Scholar] [CrossRef]
  63. Abdel-Hamid, R.M.; Bayoumi, A.; Abdellateif, M.S.; Nooh, H.A.; Refaat, L.; Kandeel, E.Z.; Hassan, S.S. Bacterial co-infections in cancer patients with COVID-19: Predictors and antimicrobial resistance trends. J. Infect. Dev. Ctries. 2024, 18, 1185–1195. [Google Scholar] [CrossRef]
  64. Gudiol, C.; Durà-Miralles, X.; Aguilar-Company, J.; Hernández-Jiménez, P.; Martínez-Cutillas, M.; Fernandez-Avilés, F.; Machado, M.; Vázquez, L.; Martín-Dávila, P.; de Castro, N.; et al. Co-infections and superinfections complicating COVID-19 in cancer patients: A multicentre, international study. J. Infect. 2021, 83, 306–313. [Google Scholar] [CrossRef]
  65. Peng, Y.; Yang, Q.; Chen, Y. Exploration of dual-acting anticancer and antimicrobial compounds: Challenges and prospects. Eur. J. Med. Chem. 2023, 266, 115089. [Google Scholar] [CrossRef]
  66. Upadhyay, A.; Jha, H.C.; Pal, D.; Kumar, A. Repurposing antibiotics: A dual-action approach against bacteria-induced cancer. Cancer Pathog. Ther. 2025, 3, E01–E11. [Google Scholar] [CrossRef]
  67. Thankappan, B.; Sivakumar, J.; Asokan, S.; Ramasamy, M.; Pillai, M.M.; Selvakumar, R.; Angayarkanni, J. Dual antimicrobial and anticancer activity of a novel synthetic α-helical antimicrobial peptide. Eur. J. Pharm. Sci. 2021, 161, 105784. [Google Scholar] [CrossRef]
  68. Elsaedany, S.K.; AbdEllatif Zein, M.; AbdElRehim, E.M.; Keshk, R.M. Synthesis and biological activity studies of novel pyrido[2,3-d]pyrimidines and pyrido[2,3-d]triazines. Russ. J. Bioorg. Chem. 2021, 47, 552–560. [Google Scholar] [CrossRef]
  69. Popiołek, Ł.; Biernasiuk, A.; Paruch, K.; Malm, A.; Wujec, M. Synthesis and in vitro antimicrobial activity screening of new pipemidic acid derivatives. Arch. Pharm. Res. 2018, 41, 633–645. [Google Scholar] [CrossRef] [PubMed]
  70. Camarasa, M.; De La Bellacasa, R.P.; González, À.L.; Ondoño, R.; Estrada, R.; Franco, S.; Badia, R.; Esté, J.; Martínez, M.Á.; Teixidó, J.; et al. Design, synthesis and biological evaluation of pyrido[2,3-d]pyrimidin-7-(8H)-ones as HCV inhibitors. Eur. J. Med. Chem. 2016, 115, 463–483. [Google Scholar] [CrossRef] [PubMed]
  71. Gao, M.; Nettles, R.E.; Belema, M.; Snyder, L.B.; Nguyen, V.N.; Fridell, R.A.; Serrano-Wu, M.H.; Langley, D.R.; Sun, J.H.; O’Boyle, D.R., II; et al. Chemical genetics strategy identifies an HCV NS5A inhibitor with a potent clinical effect. Nature 2010, 465, 96–100. [Google Scholar] [CrossRef]
  72. El-Hema, H.S.; Mady, M.F.; Abdel-Rahman, A.A.H.; Nossier, E.S.; El-Sayed, A.F.; Sabry, A.; El-Feky, M.; Hawata, M.A. Novel pyrimidopyridothiazine based derivatives as potential anticancer and antibacterial agents: A step towards design and synthesis of multi-targeted therapeutics. J. Mol. Struct. 2026, 1349, 143790. [Google Scholar] [CrossRef]
  73. Dzieszkowski, K.; Barańska, I.; Rafiński, Z. Construction of dihydropyrido[2,3-d]pyrimidine scaffolds via aza-Claisen rearrangement catalyzed by N-heterocyclic carbenes. J. Org. Chem. 2020, 85, 6645–6662. [Google Scholar] [CrossRef] [PubMed]
  74. Tantawy, E.S.; Nafie, M.S.; Morsy, H.A.; El-Sayed, H.A.; Moustafa, A.H.; Mohammed, S.M. Synthesis of novel bioactive pyrido[2,3-d]pyrimidine derivatives with potent cytotoxicity through apoptosis as PIM-1 kinase inhibitors. RSC Adv. 2024, 14, 11098–11111. [Google Scholar] [CrossRef] [PubMed]
  75. Toolabi, M.; Basiri, A.; Yaghouti, F.D.; Safdarian, M.; Ayati, A.; Mojaddami, A. Coumarin derivatives as potential anticancer agents: Synthesis, antiproliferative activity, apoptosis, and molecular docking studies. Results Chem. 2025, 16, 102442. [Google Scholar] [CrossRef]
  76. Kilic-Kurt, Z.; Ozmen, N.; Bakar-Ates, F. Synthesis and anticancer activity of some pyrimidine derivatives with aryl urea moieties as apoptosis-inducing agents. Bioorg. Chem. 2020, 101, 104028. [Google Scholar] [CrossRef]
  77. Abdelshaheed, M.M.; El Subbagh, H.I.; Tantawy, M.A.; Attia, R.T.; Youssef, K.M.; Fawzy, I.M. Discovery of new pyridine heterocyclic hybrids: Design, synthesis, dynamic simulations, and in vitro and in vivo breast cancer biological assays. RSC Adv. 2023, 13, 15689–15703. [Google Scholar] [CrossRef]
  78. Todsaporn, D.; Zubenko, A.; Kartsev, V.G.; Mahalapbutr, P.; Geronikaki, A.; Sirakanyan, S.N.; Divaeva, L.N.; Chekrisheva, V.; Yildiz, I.; Choowongkomon, K.; et al. Furopyridine derivatives as potent inhibitors of the wild type, L858R/T790M, and L858R/T790M/C797S EGFR. J. Phys. Chem. B 2024, 128, 12389–12402. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, L.; Xu, Y.; Gao, C. Underlying anti-cancer mechanisms of histone deacetylase (HDAC) inhibitors in tamoxifen-resistant breast cancer cells. Iran. J. Basic Med. Sci. 2024, 27, 775. [Google Scholar] [PubMed]
  80. Pfeffer, C.M.; Singh, A.T.K. Apoptosis: A target for anticancer therapy. Int. J. Mol. Sci. 2018, 19, 448. [Google Scholar] [CrossRef] [PubMed]
  81. Qu, Y.; Gao, C.; Li, R.; Wu, Y.; Kong, H.; Li, Y.; Li, D.; Ampomah-Wireko, M.; Wang, Y.N.; Zhang, E. Synthesis and antimicrobial evaluation of novel quaternary quinolone derivatives with low toxicity and anti-biofilm activity. Eur. J. Med. Chem. 2025, 291, 117591. [Google Scholar] [CrossRef]
  82. Radini, I.A.M. Design, synthesis, and antimicrobial evaluation of novel pyrazoles and pyrazolyl-1,3,4-thiadiazine derivatives. Molecules 2018, 23, 2092. [Google Scholar] [CrossRef]
  83. El-Hema, H.S.; Soliman, S.M.; El-Dougdoug, W.; Ahmed, M.H.; Abdelmajeid, A.; Nossier, E.S.; Hussein, M.F.; Alrayes, A.A.; Hassan, M.; Ahmed, N.A.; et al. Design, characterization, antimicrobial activity, and in silico studies of theinothiazoloquinazoline derivatives bearing thiazinone, tetrazole, and triazole moieties. ACS Omega 2025, 10, 9703–9717. [Google Scholar] [CrossRef]
  84. Mohi El-Deen, E.M.; Anwar, M.M.; El-Gwaad, A.A.A.; Karam, E.A.; El-Ashrey, M.K.; Kassab, R.R. Novel pyridothienopyrimidine derivatives: Design, synthesis and biological evaluation as antimicrobial and anticancer agents. Molecules 2022, 27, 803. [Google Scholar] [CrossRef]
  85. Hassan, S.A.; Abdullah, M.N. Synthesis of New 4-Thiazolidinone Bearing Thiazole, Assess Anticancer, and Antimicrobial Activities: Insights from DFT, and Molecular Docking. J. Mol. Struct. 2025, 1346, 143181. [Google Scholar] [CrossRef]
  86. Somkuwar, S.; Chaubey, N. In silico ADME prediction of designed pyrimidine derivatives for antiviral, antimalarial and anticancer potentials. Int. J. Pharm. Res. Dev. 2025, 7, 577–581. [Google Scholar] [CrossRef]
  87. Wu, F.; Zhou, Y.; Li, L.; Shen, X.; Chen, G.; Wang, X.; Liang, X.; Tan, M.; Huang, Z. Computational approaches in preclinical studies on drug discovery and development. Front. Chem. 2020, 8, 726. [Google Scholar] [CrossRef]
  88. Abd El-Lateef, H.M.; Ezelarab, H.A.; Ali, A.M.; Alsaggaf, A.T.; Mahdi, W.A.; Alshehri, S.; Aboelez, M.O. Design and evaluation of sulfadiazine derivatives as potent dual inhibitors of EGFR WT and EGFR T790M: Integrating biological, molecular docking, and ADMET analysis. RSC Adv. 2024, 14, 28608–28625. [Google Scholar] [CrossRef] [PubMed]
  89. Sobh, E.A.; Dahab, M.A.; Elkaeed, E.B.; Alsfouk, B.A.; Ibrahim, I.M.; Metwaly, A.M.; Eissa, I.H. New thieno[2,3-d]pyrimidine derivatives as EGFR WT and EGFR T790M inhibitors: Design, synthesis, antiproliferative activities, docking studies, ADMET, toxicity, MD simulation studies. J. Heterocycl. Chem. 2024, 61, 285–304. [Google Scholar] [CrossRef]
  90. Eissa, I.H.; Yousef, R.G.; Elkady, H.; Alsfouk, A.A.; Alsfouk, B.A.; Husein, D.Z.; Ibrahim, I.M.; Elkaeed, E.B.; Metwaly, A.M. A new anticancer semisynthetic theobromine derivative targeting EGFR protein: CADDD study. Life 2023, 13, 191. [Google Scholar] [CrossRef]
  91. Shaban, R.M.; Samir, N.; Nissan, Y.M.; Abouzid, K.A. Design, synthesis, and biological evaluation with molecular dynamics study of novel pyrazolo[3,4-d]pyrimidine derivatives as anti-cancer agents. RSC Adv. 2023, 13, 17074–17096. [Google Scholar] [CrossRef]
  92. Aboelez, M.O.; Belal, A.; Xiang, G.; Ma, X. Design, synthesis, and molecular docking studies of novel pomalidomide-based PROTACs as potential anti-cancer agents targeting EGFR WT and EGFR T790M. J. Enzyme Inhib. Med. Chem. 2022, 37, 1196–1211. [Google Scholar] [CrossRef] [PubMed]
  93. Zhou, W.; Ercan, D.; Chen, L.; Yun, C.H.; Li, D.; Capelletti, M.; Cortot, A.B.; Chirieac, L.; Iacob, R.E.; Padera, R.; et al. Novel mutant-selective EGFR kinase inhibitors against EGFR T790M. Nature 2009, 462, 1070–1074. [Google Scholar] [CrossRef] [PubMed]
  94. Holdgate, G.A.; Tunnicliffe, A.; Ward, W.H.; Weston, S.A.; Rosenbrock, G.; Barth, P.T.; Taylor, I.W.; Pauptit, R.A.; Timms, D. The entropic penalty of ordered water accounts for weaker binding of the antibiotic novobiocin to a resistant mutant of DNA gyrase: A thermodynamic and crystallographic study. Biochemistry 1997, 36, 9663–9673. [Google Scholar] [CrossRef] [PubMed]
  95. Love, R.A.; Brodsky, O.; Hickey, M.J.; Wells, P.A.; Cronin, C.N. Crystal structure of a novel dimeric form of NS5A domain I protein from hepatitis C virus. J. Virol. 2009, 83, 4395–4403. [Google Scholar] [CrossRef]
  96. Abo-Salem, H.M.; Ali, E.A.; El-Mowafi, S.A.; Abdel-Aziz, M.S.; El-Sawy, E.R.; Abd El Salam, H.A. New sulfonamide-tethered coumarin derivatives as potential DNA gyrase inhibitors: Design, synthesis, antimicrobial evaluation, and in silico study. J. Mol. Struct. 2024, 1296, 136860. [Google Scholar] [CrossRef]
Figure 1. Anticancer agents consist of erlotinib as a reference I and other pyrido[2,3-d] pyrimidines IIVII that inhibit EGFR and its mutations.
Figure 1. Anticancer agents consist of erlotinib as a reference I and other pyrido[2,3-d] pyrimidines IIVII that inhibit EGFR and its mutations.
Pharmaceuticals 18 01472 g001
Figure 2. Pyrido[2,3-d]pyrimidine-based analogues VIIIXI bearing antimicrobial and antiviral activity.
Figure 2. Pyrido[2,3-d]pyrimidine-based analogues VIIIXI bearing antimicrobial and antiviral activity.
Pharmaceuticals 18 01472 g002
Figure 3. The rational design of newly synthesized 2,4-dioxopyrido[2,3-d]pyrimidines 116 with expected anticancer and antimicrobial activities through inhibition of EGFR and DNA gyrase, respectively.
Figure 3. The rational design of newly synthesized 2,4-dioxopyrido[2,3-d]pyrimidines 116 with expected anticancer and antimicrobial activities through inhibition of EGFR and DNA gyrase, respectively.
Pharmaceuticals 18 01472 g003
Scheme 1. Synthesis of Pyridopyrimidine, acid hydrazide, hydrazinyl pyrazole, amino-triazole-3-thiol, and thio hydrazide derivatives 15.
Scheme 1. Synthesis of Pyridopyrimidine, acid hydrazide, hydrazinyl pyrazole, amino-triazole-3-thiol, and thio hydrazide derivatives 15.
Pharmaceuticals 18 01472 sch001
Scheme 2. Synthesis of dimethyl-pyrazole, oxadiazole-2-thiol, and amino-pyrazole-4-carbonitrile derivatives 68.
Scheme 2. Synthesis of dimethyl-pyrazole, oxadiazole-2-thiol, and amino-pyrazole-4-carbonitrile derivatives 68.
Pharmaceuticals 18 01472 sch002
Scheme 3. Synthesis of thioxo-oxadiazol-propanenitrile and ethyl tetrazole derivatives 9 and 10.
Scheme 3. Synthesis of thioxo-oxadiazol-propanenitrile and ethyl tetrazole derivatives 9 and 10.
Pharmaceuticals 18 01472 sch003
Scheme 4. Synthesis of fused -pyrazolopyrimidin-4-one and pyrazolopyrimidin-4-amine derivatives 11 and 12.
Scheme 4. Synthesis of fused -pyrazolopyrimidin-4-one and pyrazolopyrimidin-4-amine derivatives 11 and 12.
Pharmaceuticals 18 01472 sch004
Scheme 5. Synthesis of dimethyl pyrazole derivatives 13.
Scheme 5. Synthesis of dimethyl pyrazole derivatives 13.
Pharmaceuticals 18 01472 sch005
Scheme 6. Chloro triazolothiadiazine and triazolothiadiazole-6-thiol derivatives 14 and 15 synthesis.
Scheme 6. Chloro triazolothiadiazine and triazolothiadiazole-6-thiol derivatives 14 and 15 synthesis.
Pharmaceuticals 18 01472 sch006
Scheme 7. Synthesis of methyl-pyrazol-3-one derivative 16.
Scheme 7. Synthesis of methyl-pyrazol-3-one derivative 16.
Pharmaceuticals 18 01472 sch007
Figure 4. The cytotoxic effects of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 116 on WI-38 against HeLa, HepG-2, and MCF-7 cell lines that was tested at multiple doses employing the MTT assay, represented by IC50 values.
Figure 4. The cytotoxic effects of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 116 on WI-38 against HeLa, HepG-2, and MCF-7 cell lines that was tested at multiple doses employing the MTT assay, represented by IC50 values.
Pharmaceuticals 18 01472 g004
Figure 5. SAR graphic of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 116 that were identified as possible cytotoxic agents targeting the human malignant cells HeLa, HepG-2, and MCF-7. Ar = aromatic group; R = aliphatic group.
Figure 5. SAR graphic of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 116 that were identified as possible cytotoxic agents targeting the human malignant cells HeLa, HepG-2, and MCF-7. Ar = aromatic group; R = aliphatic group.
Pharmaceuticals 18 01472 g005
Figure 6. Effect of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivative (Compound 1) on the percentage of MCF-7 cells showing V-FITC-positive annexin staining, along with cell cycle analysis.
Figure 6. Effect of 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivative (Compound 1) on the percentage of MCF-7 cells showing V-FITC-positive annexin staining, along with cell cycle analysis.
Pharmaceuticals 18 01472 g006
Figure 7. Cell cycle analysis of Compound 1 (1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivative).
Figure 7. Cell cycle analysis of Compound 1 (1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivative).
Pharmaceuticals 18 01472 g007
Figure 8. 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1’s impact upon apoptotic activity.
Figure 8. 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1’s impact upon apoptotic activity.
Pharmaceuticals 18 01472 g008
Figure 9. Wound healing assay for MCF-7 cells treated with 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1 compared with control untreated cells.
Figure 9. Wound healing assay for MCF-7 cells treated with 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1 compared with control untreated cells.
Pharmaceuticals 18 01472 g009
Figure 10. The antibacterial activity of samples towards S. aureus ATCC25923, E. faecalis ATCC29212, P. aeruginosa ATCC10145, and E. coli ATCC25915, as determined by the well diffusion method.
Figure 10. The antibacterial activity of samples towards S. aureus ATCC25923, E. faecalis ATCC29212, P. aeruginosa ATCC10145, and E. coli ATCC25915, as determined by the well diffusion method.
Pharmaceuticals 18 01472 g010
Figure 11. (A) Bioavailability radar plots of compounds 1 and 2 showing predicted (red) and optimal (pink) values. (B) Boiled egg diagram illustrating gastrointestinal absorption, blood–brain barrier penetration, and P-glycoprotein substrate status (GP− = non-substrate, GP+ = substrate).
Figure 11. (A) Bioavailability radar plots of compounds 1 and 2 showing predicted (red) and optimal (pink) values. (B) Boiled egg diagram illustrating gastrointestinal absorption, blood–brain barrier penetration, and P-glycoprotein substrate status (GP− = non-substrate, GP+ = substrate).
Pharmaceuticals 18 01472 g011
Figure 12. Panels (A) and (B) present the 2D and 3D binding orientations of the active 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives 1 and 2 inside the binding pocket of wild-type EGFR (PDB ID: 1M17), respectively.
Figure 12. Panels (A) and (B) present the 2D and 3D binding orientations of the active 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives 1 and 2 inside the binding pocket of wild-type EGFR (PDB ID: 1M17), respectively.
Pharmaceuticals 18 01472 g012
Figure 13. Diagrams (A,B) illustrate the 2D and 3D conformations of the active 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives 1 and 2 within the EGFRT790M active site (PDB code: 3IKA), respectively.
Figure 13. Diagrams (A,B) illustrate the 2D and 3D conformations of the active 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine derivatives 1 and 2 within the EGFRT790M active site (PDB code: 3IKA), respectively.
Pharmaceuticals 18 01472 g013
Figure 14. Two and three-dimensional views of promising 1,3-dimethyl-2,4-dioxopyrid [2,3-d]pyrimidines 1 within the E. coli DNA gyrase active site (PDB code: 1AJ6).
Figure 14. Two and three-dimensional views of promising 1,3-dimethyl-2,4-dioxopyrid [2,3-d]pyrimidines 1 within the E. coli DNA gyrase active site (PDB code: 1AJ6).
Pharmaceuticals 18 01472 g014
Figure 15. Two and three-dimensional views of promising 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 1 within the HCV NS5A active site (PDB code: 3FQM).
Figure 15. Two and three-dimensional views of promising 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 1 within the HCV NS5A active site (PDB code: 3FQM).
Pharmaceuticals 18 01472 g015
Figure 16. RMSD assessments for the MD simulations are shown in (A) and for 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1 with EGFRWT, and mutant EGFRT790M in (B) respectively.
Figure 16. RMSD assessments for the MD simulations are shown in (A) and for 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1 with EGFRWT, and mutant EGFRT790M in (B) respectively.
Pharmaceuticals 18 01472 g016
Figure 17. Molecular dynamics trajectory graphs (A,B) for 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1 with EGFRWT, and mutant EGFRT790M, respectively.
Figure 17. Molecular dynamics trajectory graphs (A,B) for 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine 1 with EGFRWT, and mutant EGFRT790M, respectively.
Pharmaceuticals 18 01472 g017
Figure 18. Trajectories (A,B) illustrate the interaction patterns of compound 1 (1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine) with EGFRWT and the T790M mutant, respectively.
Figure 18. Trajectories (A,B) illustrate the interaction patterns of compound 1 (1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine) with EGFRWT and the T790M mutant, respectively.
Pharmaceuticals 18 01472 g018
Figure 19. Timelines (A,B) illustrate the interaction and contact profiles of compound 1 (1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine) with EGFRWT and the T790M mutant, respectively.
Figure 19. Timelines (A,B) illustrate the interaction and contact profiles of compound 1 (1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidine) with EGFRWT and the T790M mutant, respectively.
Pharmaceuticals 18 01472 g019
Figure 20. DFT-calculated geometries and frontier orbital (HOMO/LUMO) maps of compounds 1 and 2.
Figure 20. DFT-calculated geometries and frontier orbital (HOMO/LUMO) maps of compounds 1 and 2.
Pharmaceuticals 18 01472 g020aPharmaceuticals 18 01472 g020b
Table 1. IC50 Inhibitory impact (µM) in vitro of the potentially valuable 1,3-dimethyl uracil 1 and 2 against wild EGFR, mutant EGFR (L858R), and EGFR (T790M), compared with erlotinib and fold-change in potency relative to wild-type EGFR, illustrating resistance impact.
Table 1. IC50 Inhibitory impact (µM) in vitro of the potentially valuable 1,3-dimethyl uracil 1 and 2 against wild EGFR, mutant EGFR (L858R), and EGFR (T790M), compared with erlotinib and fold-change in potency relative to wild-type EGFR, illustrating resistance impact.
Compd. No.IC50 (Mean ± SD) (µM)Fold-Change
EGFREGFR_L858REGFR_T790ML858R/WTT790M/WT
10.093 ± 0.0031.157 ± 0.0360.174 ± 0.00612.44×1.87×
20.212 ± 0.0071.603 ± 0.0500.854 ± 0.0277.56×4.03×
Erlotinib0.051 ± 0.0020.239 ± 0.0070.098 ± 0.0034.69×1.92×
SD reflects variability from three measurements, × expresses fold change, while IC50 is defined as the concentration necessary to achieve 50% enzyme inhibition.
Table 2. Agar diffusion technique MIC of the most active chemicals.
Table 2. Agar diffusion technique MIC of the most active chemicals.
CompoundsMIC (µg/mL)
E. coli
ATCC25915
S. aureus
ATCC25923
E. faecalis ATCC29212 P. aeruginosa
ATCC10145
150 10010050
1250 >10010050
1350 >100>100100
1650 1001005
Table 3. Physicochemical properties calculations of the promising 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 1 and 2.
Table 3. Physicochemical properties calculations of the promising 1,3-dimethyl-2,4-dioxopyrido[2,3-d]pyrimidines 1 and 2.
Compd.Violations aMW bnHBD cnHBA dnRB eTPSA (Å2) fMLogP g
Rule-≤500≤5≤10≤10≤140≤4.15
10 (Lipinski & Veber)388.80154109.212.86
20 (Lipinski & Veber)374.78 353138.031.34
a Violations from Lipinski and Veber Rules; b Molecular Weight; c Number of Hydrogen Bond Donor; d Number of Hydrogen Bond Acceptor; e Number of Rotatable Bond; f Calculated Lipophilicity (MLog Po/w); g Topological Polar Surface Area.
Table 4. DFT-Based Quantum Chemical Calculations and Reactivity Descriptors of synthesized Compounds.
Table 4. DFT-Based Quantum Chemical Calculations and Reactivity Descriptors of synthesized Compounds.
Cpds.LUMOHOMOΔEAIXηSωΔN maxΔN
1−0.168−0.27440.10640.1680.27440.22120.053218.796990.9197262.07894763.71053
2−0.17313−0.177910.004780.173130.177910.175520.00239418.4112.8900736.719671427.715
3−0.17124−0.216470.045230.171240.216470.1938550.02261544.218441.6617184.285983150.4786
4−0.18425−0.241450.05720.184250.241450.212850.028634.965031.5840953.721154118.6565
5−0.1689−0.227210.058310.16890.227210.1980550.02915534.299431.3454223.396587116.6514
6−0.21131−0.250650.039340.211310.250650.230980.0196750.838842.7123425.871378172.0646
7−0.16429−0.205770.041480.164290.205770.185030.0207448.216011.6507284.460704164.2953
8−0.17823−0.208280.030050.178230.208280.1932550.01502566.555742.485696.431115226.514
9−0.16436−0.202010.037650.164360.202010.1831850.01882553.120851.7825634.865471181.0575
10−0.16012−0.239050.078930.160120.239050.1995850.03946525.338911.0093542.52863386.15754
11−0.16261−0.191050.028440.162610.191050.176830.0142270.323492.1989356.217651239.9146
12−0.17746−0.255030.077570.177460.255030.2162450.03878525.783161.205672.7877487.45333
13−0.17528−0.216910.041630.175280.216910.1960950.02081548.042281.8473824.710425163.4375
14−0.18373−0.214690.030960.183730.214690.199210.0154864.599482.5636066.434432219.6638
15−0.16836−0.204040.035680.168360.204040.18620.0178456.053811.943415.21861190.9697
16−0.18343−0.22660.043170.183430.22660.2050150.02158546.328471.9472394.749016157.4006
Table 5. Quantum Chemical Descriptors and ESP-Based Composite Bioactivity Prediction of Synthesized Compounds.
Table 5. Quantum Chemical Descriptors and ESP-Based Composite Bioactivity Prediction of Synthesized Compounds.
CompoundΔESωESP ScoreComposite Bioactivity Score
10.106418.796990.9197260.1879370.636331
20.00478418.4112.890070.8707480.830612
30.0452344.218441.6617180.2450070.188622
40.057234.965031.5840950.1879370.637946
50.0583134.299431.3454220.3789231.083423
60.0393450.838842.7123420.2525310.223867
70.0414848.216011.6507280.2310740.175793
80.0300566.555742.485690.2158030.244654
90.0376553.120851.7825630.1185380.261083
100.0789325.338911.0093540.1572920.169537
110.0284470.323492.1989350.51490.290725
120.0775725.783161.205670.151.066763
130.0416348.042281.8473820.2179970.326359
140.0309664.599482.5636060.1844590.237995
150.0356856.053811.943410.2955710.275627
160.0431746.328471.9472390.2727970.336711
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

El-Hema, H.S.; Shehata, H.E.; Hawata, M.A.; Nossier, E.S.; El-Sayed, A.F.; Altwaijry, N.A.; Saleh, A.; Hussein, M.F.; Sabry, A.; Abdel-Rahman, A.A.-H. Innovative Amino-Functionalization of Pyrido[2,3-d]pyrimidine Scaffolds for Broad Therapeutic Applications Supported by Computational Analyses. Pharmaceuticals 2025, 18, 1472. https://doi.org/10.3390/ph18101472

AMA Style

El-Hema HS, Shehata HE, Hawata MA, Nossier ES, El-Sayed AF, Altwaijry NA, Saleh A, Hussein MF, Sabry A, Abdel-Rahman AA-H. Innovative Amino-Functionalization of Pyrido[2,3-d]pyrimidine Scaffolds for Broad Therapeutic Applications Supported by Computational Analyses. Pharmaceuticals. 2025; 18(10):1472. https://doi.org/10.3390/ph18101472

Chicago/Turabian Style

El-Hema, Hagar S., Haitham E. Shehata, Mohamed A. Hawata, Eman S. Nossier, Ahmed F. El-Sayed, Najla A. Altwaijry, Asmaa Saleh, Modather F. Hussein, Amr Sabry, and Adel A.-H. Abdel-Rahman. 2025. "Innovative Amino-Functionalization of Pyrido[2,3-d]pyrimidine Scaffolds for Broad Therapeutic Applications Supported by Computational Analyses" Pharmaceuticals 18, no. 10: 1472. https://doi.org/10.3390/ph18101472

APA Style

El-Hema, H. S., Shehata, H. E., Hawata, M. A., Nossier, E. S., El-Sayed, A. F., Altwaijry, N. A., Saleh, A., Hussein, M. F., Sabry, A., & Abdel-Rahman, A. A.-H. (2025). Innovative Amino-Functionalization of Pyrido[2,3-d]pyrimidine Scaffolds for Broad Therapeutic Applications Supported by Computational Analyses. Pharmaceuticals, 18(10), 1472. https://doi.org/10.3390/ph18101472

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop