Next Article in Journal
2-Hydroxyestradiol Overcomes Mesenchymal Stem Cells-Mediated Platinum Chemoresistance in Ovarian Cancer Cells in an ERK-Independent Fashion
Next Article in Special Issue
Exploring the Synergistic Anticancer Potential of Benzofuran–Oxadiazoles and Triazoles: Improved Ultrasound- and Microwave-Assisted Synthesis, Molecular Docking, Hemolytic, Thrombolytic and Anticancer Evaluation of Furan-Based Molecules
Previous Article in Journal
Understanding Parahydrogen Hyperpolarized Urine Spectra: The Case of Adenosine Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Pyridothienopyrimidine Derivatives: Design, Synthesis and Biological Evaluation as Antimicrobial and Anticancer Agents

by
Eman M. Mohi El-Deen
1,*,
Manal M. Anwar
1,
Amina A. Abd El-Gwaad
1,
Eman A. Karam
2,
Mohamed K. El-Ashrey
3 and
Rafika R. Kassab
4
1
Department of Therapeutic Chemistry, National Research Centre, Cairo 12622, Egypt
2
Department of Microbial Chemistry, National Research Centre, Cairo 12622, Egypt
3
Department of Pharmaceutical Chemistry, Faculty of Pharmacy, Cairo University, Cairo 11562, Egypt
4
Department of Chemistry, Faculty of Science (Girls), Al-Azhar University, Cairo 11754, Egypt
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(3), 803; https://doi.org/10.3390/molecules27030803
Submission received: 30 December 2021 / Revised: 14 January 2022 / Accepted: 19 January 2022 / Published: 26 January 2022
(This article belongs to the Special Issue Heterocyclic Compounds: Design, Synthesis, and Applications)

Abstract

:
The growing risk of antimicrobial resistance besides the continuous increase in the number of cancer patients represents a great threat to global health, which requires intensified efforts to discover new bioactive compounds to use as antimicrobial and anticancer agents. Thus, a new set of pyridothienopyrimidine derivatives 2a,b–9a,b was synthesized via cyclization reactions of 3-amino-thieno[2,3-b]pyridine-2-carboxamides 1a,b with different reagents. All new compounds were evaluated against five bacterial and five fungal strains. Many of the target compounds showed significant antimicrobial activity. In addition, the new derivatives were further subjected to cytotoxicity evaluation against HepG-2 and MCF-7 cancer cell lines. The most potent cytotoxic candidates (3a, 4a, 5a, 6b, 8b and 9b) were examined as EGFR kinase inhibitors. Molecular docking study was also performed to explore the binding modes of these derivatives at the active site of EGFR-PK. Compounds 3a, 5a and 9b displayed broad spectrum antimicrobial activity with MIC ranges of 4–16 µg/mL and potent cytotoxic activity with IC50 ranges of 1.17–2.79 µM. In addition, they provided suppressing activity against EGFR with IC50 ranges of 7.27–17.29 nM, higher than that of erlotinib, IC50 = 27.01 nM.

1. Introduction

In the past twenty years, the incidence of microbial infections associated with increased antimicrobial resistance has increased by alarming levels worldwide, endangering the possibility of curing many infectious diseases, and representing a global threat to population health [1,2,3]. One of the potential approaches to combat this resistance problem is based on designing innovative new molecules with different modes of action to overcome cross-resistance to present therapeutics [4,5]. Efforts in developing broad-spectrum as well as specific and targeted drugs against various microbes are continuous in all directions [5]. One of the recent strategies used for developing new antimicrobial agents is hybridization of different pharmacophores binding diverse biomolecular targets to exert synergistic effects against drug-resistant infectious diseases [6].
On the other hand, despite large advances in the diagnosis and treatment of various types of cancers, the survival of cancer patients remains poor because of the widespread side effects of anticancer therapeutics as well as the acquisition of multiple drug resistance by the cancer cells [7]. Breast and liver cancers are among the most medically significant cancers due to their high incidence and morbidity [8]. Therefore, obtaining new, effective, more selective, and less toxic anticancer agents remains one of the most urgent demands [9]. In addition, cancer patients are at higher risk of drug-resistant infectious diseases because of the weakness and immunosuppression caused by anticancer drug therapy [10], which necessitates the development of new drugs that have potent dual activity against cancer and pathogenic microbes [11].
Heterocyclic ring systems are considered key molecular structures in medicinal chemistry. In addition to their presence in the skeleton of many biological molecules such as hemoglobin, DNA, RNA, vitamins, hormones, and heterocyclic compounds [12,13,14], they produce a wide range of biological activities [15,16]. Accordingly, the structural diversity and biological prominence of heterocyclic compounds have made them attractive synthetic targets in drug design and discovery for many years [12,13]. Thieno[2,3-b]pyridines constitute a set of heterocyclic compounds that have beneficial effects in the treatment of many diseases. Recently, many studies have described various substituted thienopyridine compounds as significant antiproliferative agents against a wide range of human cancer cell lines [17,18,19] and as antimicrobial [20,21,22], anti-Alzheimer’s [23], anti-platelet [24], antiviral [25,26] and anti-inflammatory [27] candidates.
Furthermore, the pyrimidine structure is closely related to the nucleobases—uracil, thymine, and cytosine—which makes pyrimidine molecules important building blocks in all living cells [28,29]. Pyrimidine-based compounds exhibit a broad spectrum of pharmacological activity, such as antimalarial [30], antidiabetic [31], antimicrobial [32,33], and anticancer activities [34,35,36]. Therefore, combining thienopyridine and pyrimidine cores in the same molecular architecture, forming a pyridothienopyrimidine nucleus, serves as an attractive strategy for designing a novel scaffold with more favorable pharmacological effects [37]. Recently, various pyridothienopyrimidine derivatives were reported to produce significant antimicrobial [38,39] and anticancer activities [40,41], as well as to suppress protein kinases such as serine/threonine kinase and vascular endothelial growth factor receptor (VEGFR-2) [42,43].
Prompted by the abovementioned issues, the strategy of this work was focused on the design of new tricyclic pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine compounds with structures that hybridize features of the reported thienopyridine and pyrimidine antimicrobial and/or anticancer agents [21,22,33,34] Figure 1. The resulting new pyridothienopyrimidines 2a,b–9a,b were synthesized via different cyclization reactions of the key starting compounds, 3-aminothieno[2,3-b]pyridine-2-carboxamides 1a,b. The tricyclic ring system of the target pyridothienopyridine compounds was linked at position-2 and/or position-4 to different privileged structural motifs such as morpholine, piprazine, spiro cycloalkane, oxirane, and acetamide, which are renowned for their valuable and diverse pharmacological activity [44,45,46,47,48]. All synthesized derivatives were evaluated as antimicrobial agents against a panel of Gram-positive and Gram-negative bacterial and fungal pathogens. Then, they were evaluated as cytotoxic agents against human liver carcinoma cells (HepG2) and human breast cancer cells (MCF-7) in an effort to gain new compounds possessing dual antimicrobial and anticancer activities. Since epidermal growth factor receptor (EGFR) is a key mediator in the regulation of different important cellular processes [49,50] and its overexpression displays a significant role in the development of many human cancers, including hepatic cancer and breast cancer [51,52], it was of interest to examine the EGFR inhibition activities of the compounds that revealed the most potent cytotoxic activities, as one of their cytotoxic mechanisms of action. Furthermore, molecular docking study was carried out to detect the compounds’ binding modes at the active site of EGFR-TK.

2. Results and Discussion

2.1. Chemistry

The synthetic pathways utilized for the synthesis of the target pyridothienopyrimidine compounds 2a,b9a,b are depicted in Scheme 1 and Scheme 2. The structures of all new compounds were confirmed using 1H-NMR, 13C-NMR (Supplementary Materials), IR, and mass spectral data alongside the elemental microanalyses. Heating 3-amino-4,6-dimethyl/6-phenyl-4-(p-tolyl)thieno[2,3-b]pyridine-2-carboxamides 1a,b, the starting compounds, with urea at 180 °C for 1 h led to the formation of the corresponding pyridothienopyrimidine-2,4-diones 2a,b, which were further refluxed with a mixture of phosphorus oxychloride and phosphorus pentachloride to give the corresponding 2,4-dichloro derivatives 3a,b, respectively. On the other hand, the treatment of the starting compounds 1a,b with 2-chloroacetyl chloride in cold acetonitrile resulted in cyclization via substitution reaction followed by intramolecular elimination to give the 2-(chloromethyl)-pyridothienopyrimidin-4(3H)-ones 4a,b, respectively. The subsequent treatment of 4b with different amines, namely, morpholine and 1-methylpiperazine, resulted in the corresponding 2-(morpholinomethyl)-/2-((4-methylpiperazin-1-yl)methyl)-pyridothieopyrimidin-4(3H)-ones 5a,b (Scheme 1). IR spectra of compounds 2a,b found two bands in the region 3394–3112 cm−1 ascribed to 2NH and two bands in the region 1728–1640 cm−1 related to the 2C=O groups of the pyrimidine-2,4(1H,3H)-dione ring. By comparison, IR spectra of the 2,4-dichloro derivatives 3a,b revealed the absence of the previous bands; instead, they showed two bands in the region 825–768 cm−1 referring to the newly formed C-Cl groups. In addition, 1H-NMR spectra of compounds 2a,b revealed the two 2NH groups to be two D2O exchangeable singlets at δ 10.42–11.71 ppm, which vanished in the 1H-NMR spectra of 2,4-dichloro derivatives 3a,b. Furthermore, the 2CH3 of 2a, 3a and CH3 of the p-tolyl moiety of 2b, 3b exhibited as singlet signals in the range 2.43–2.79 ppm along with aromatic protons in the range δ 7.21–8.24 ppm. The 13C NMR spectra of compounds 2a,b and 3a,b revealed CH3 moieties at δ 20.40–24.36 ppm in addition to the aromatic carbons.
IR spectra of compounds 4a,b, 5a,b represented NH and C=O groups as absorption bands in the regions 3440–3387 cm−1 and 1669–1655 cm−1, respectively, while the C-Cl band of 4a appeared at 772 cm−1 and that of 4b was at 776 cm−1. Furthermore, 1H-NMR spectra of 4a,b and 5a,b exhibited, in addition to the signals of the parent protons, a singlet signal in the region δ 3.29–4.66, referring to the methylene protons of the -NCH2 moiety. Additionally, the eight protons of the morpholine moiety of 5a appeared as two singlets at δ 2.38 and 3.54 ppm, while those of the piperazine ring of 5b appeared as multiplet signals in the range δ 2.81–3.01 ppm alongside a singlet signal at δ 2.54 ppm due to its -N-CH3 group. The 13C-NMR spectra of 5a exhibited the 2CH2N- and 2CH2O of the morpholine moiety as two signals at δ 53.37 and 60.65 ppm, respectively.
The synthesis of spiro cycloalkane-pyridothienopyrimidin-4′-ones 6a–d in good yields was achieved via cyclocondensation reactions of the starting 1a,b with the appropriate cyclic ketones, cyclopentanone and/or cyclohexanone, in DMF containing zinc chloride anhydrous. Treating the latter compounds (6a–d) with phosphorus pentasulfide in refluxing pyridine yielded the corresponding 1′H-spiro[cycloalkane-1,2′-pyridothienopyrimidine]-4′(3′H)-thiones 7a–d (Scheme 2). IR spectra of 6a–d and 7a–d revealed the presence of different absorption bands in the regions 3427–3157 cm−1 related to 2NH groups, 1660–1644 cm−1 related to the C=O groups of 6a-d and 1247–1232 cm−1 related to the C=S groups of compounds 7a–d. 1H-NMR spectra of 6a–d and 7a–d represented the eight protons of the spiro-cyclopentane and the ten protons of the spiro-cyclohexane substituents as multiplet signals in the range δ 1.21–2.11 ppm alongside one D2O exchangeable singlet in the range δ 4.59–6.49 ppm, corresponding to the NH at position-1, and another D2O exchangeable singlet appeared downfield in the range δ 7.81–9.85 ppm, related to the NH group at position-3. 13C-NMR spectra of 6a–d and 7a–d revealed various singlets in the ranges δ 19.26–38.33 ppm assignable to the cyclopentane and cyclohexane carbons, δ 69.49–70.81 ppm related to the spiro carbons, δ 164.40–167.80 ppm assigned to the C=O groups of 6a–d and δ 181.33–183.35 ppm ascribed to C=S moieties of compounds 7a–d, as well as the signals of the parent carbons, which appeared in their correct regions.
Compounds 7c,b were treated with 2-chloroacetamide in refluxing DMF containing sodium carbonate anhydrous to obtain the corresponding 4-sulfanyl acetamide derivatives 8a,b, while the treatment of 7c,b with epichlorohydrin in refluxing acetone containing a catalytic amount of triethyl amine resulted in the formation of the corresponding 4-(oxiran-2-ylmethy)sulfanyl derivatives 9a,b (Scheme 2). The S-alkylation at position-4 was supported by 1H NMR and 13C-NMR spectra of 8a,b and 9a,b due to the vanishing of both the signal related to one of the two NH protons and that of the C=S carbon. Moreover, the 1H NMR spectra of compounds 8a,b exhibited singlet signals at δ 4.05 and 4.19 ppm due to SCH2 protons of the newly formed acetamide side chain. 1H NMR spectra of compounds 9a,b represented the protons of the oxirane ring as two multiplets in the range δ 2.86–2.94 ppm, related to the methylene protons, and a third multiplet at δ 3.72–3.95 ppm, due to the methine proton, while the methylene protons of SCH2 appeared as a doublet signal at δ 3.51 and 3.67 ppm. Additionally, the 13C-NMR spectra 9a,b showed three signals at the range δ 31.88–53.45 ppm referred to SCH2, OCH2, and OCH moieties.
Confirmation of the molecular structures of the new compounds was also supported by their mass spectra, which represented the correct molecular ion peaks.

2.2. Biological Activity

2.2.1. Antimicrobial Activity

The newly synthesized pyridothienopyrimidine componds 2a,b–9a,b were investigated for their antimicrobial activities against a panel of microbial strains, three Gram-positive bacteria viz. Staphylococcus aureus 25923, Bacillus subtilis 6633, Bacillus cereus 33018, two Gram-negative bacteria viz. Escherichia coli 8739, Salmonella typhimurium 14028, three yeasts viz. Candida albicans 10231, Candida tropicals 750, Saccharomycese cerevisiae and two fungi viz. Aspergillus flavus, Aspergillus niger EM77 (KF774181). Their activities were expressed in terms of minimal inhibitory concentration (MIC) (μg/mL). Additionally, amoxicillin trihydrate and clotrimazole were utilized as positive antibiotic and antifungal controls. The obtained MIC results are represented in Table 1 and Table 2 and Figure 2 and Figure 3.
Based on the MIC values in Table 1, the antibacterial activity of the target compounds 2a,b9a,b was compared with that of amoxicillin, whose MICs ranged between 4–16 μg/mL against the five bacterial strains. It is evident that pyridothienopyrimidine-2,4(1H,3H)-dione derivatives 2a,b showed antibacterial activity varying from weak to inactive against the tested strains, with MICs ranging from 64 to >128 μg/mL. However, the conversion of 2a,b to 2,4-dichloro derivatives 3a,b led to a significant increase in antibacterial activity, especially for 7,9-dimethyl derivative 3a, which exhibited more potent activity than amoxicillin against B. subtilis and B. cereus and equalized with it against the other strains. In addition, 2-chloromethyl derivatives 4a,b revealed significant antibacterial activity, particularly 7,9-dimethyl derivative 4a, which showed the most potent activity against B. cereus with MIC = 4 μg/mL and had equipotent activity to amoxicillin against the other strains. While 7-Phenyl-9-(p-tolyl) analogue 4b revealed potent activity against S. aureus and B. cereus, it showed weak activity against other bacterial strains with MIC = 128 μg/mL. Further reaction of 2-chloromethyl derivative 4b with amines led to enhanced antibacterial activity, where 2-morpholinomethyl derivative 5a showed potent activity equal to amoxicillin against all the tested bacterial strains except S. aureus, with MICs ranging from 8 to 16 μg/mL. In addition, 2-(4-methylpiperazin-1-yl) derivative 5b revealed increasing activity against B. subtilis, E. coli and S. typhimurium.
Spiro cycloalkane-1,2′-pyridothienopyrimidin]-4′(3′H)-ones 6a–d exhibited antibacterial activity varying from potent to moderate against the tested strains, with MICs ranging from 4 to 32 μg/mL. Compound 6b showed an activity equal to that of the reference drug against the five bacterial strains, while 6c exceeded the potency of amoxicillin against B. subtilis and B. cereus and gave equipotent activity against the other strains. The other derivatives, 6a and 6d, showed activity varying from potent to moderate with MICs ranging 8–32 μg/mL. The conversion of 6a–d to 4′(3′H)-thiones analogues 7a–d resulted in an obvious weakening in the activity against the most of the tested strains, particularly derivative 7b, which showed weak activity against the five strains with MICs in the range 64–128 μg/mL. However, the S-alkylation of 7b and 7c at position-4 afforded increases in the antibacterial activity. The spiro cyclopentane 4-sulfanylacetamide derivative 8b revealed potent activity similar to that of amoxicillin, and the spiro cyclohexane analogue 8a showed potent to moderate activity with MICs ranging from 8 to 32 μg/mL. Spiro cyclopentane 4-(oxiran-2-ylmethyl)sulfanyl derivative 9b showed the most potent antibacterial activity against all tested strains, especially against E. coli, while spiro cyclohexane analogue 9a showed antibacterial activity ranging from potent to moderate with MICs in the range 8–32 μg/mL.
The antifungal activity of the new compounds 2a,b9a,b was evaluated according to their MIC values in comparison to the MIC values of the reference antifungal drug clotrimazole, listed in Table 2. The target compounds (3a, 4a, 4b, 5a, 6b, 6c, 7a, 7c, 8b and 9b) appeared to be potent antifungal candidates, representing MIC values ranging from 4 to 16 μg/mL against all the tested yeasts and fungi strains, similar to the range of clotrimazole (MICs; 4–16 μg/mL). Furthermore, 4-(oxiran-2-ylmethyl)sulfanyl derivative 9b represented more potent antifungal activity than that of clotrimazole against the yeast pathogens C. albicans and C. tropicals, with MICs of 4 and 4 μg/mL, respectively. The rest of the target compounds displayed moderate to weak activity against the tested yeasts and fungi strains. The obtained results suggested that the conjugation of chloroine, chloromethyl, morpholinomethyl and spiro cyclopentane/cyclohexane at position-2 of the parent pyridothienopyrimid-4(3H)-one (3a, 4a,b, 5a, 6b,c, 7c) as well as the attachment of (oxiran-2-ylmethyl)sulfanyl and/or sulfanylacetamide side chains at position-4 of the pyridothienopyrimidine scaffold (8b, 9b) produced a beneficial impact for antimicrobial activity, resulting in promising broad-spectrum growth inhibition activity against the examined Gram-positive and Gram-negative microbes as well as fungal pathogens.

2.2.2. Cytotoxic Activity

The newly synthesized pyridothienopyrimidine derivatives 2a,b9a,b were subjected to antiproliferative activity evaluation against two cancer cell lines, hepatocellular carcinoma (HepG2) and breast cancer (MCF-7), by using the MTT colorimetric assay[53]. The cytotoxic activity of the target compounds was compared with that of doxorubicin, utilized as a positive control. The concentrations of the examined derivatives that induced 50% inhibition of cell viability (IC50, μM) were detected and are listed in Table 3.
Based on IC50 values from Table 3, the examined compounds displayed versatile anti-proliferative activities against the tested cell lines, producing more potent growth inhibitory activity against HepG2 cells than MCF-7 cells with IC50 ranges of 1.17–56.18 µM and 1.52–77.41μM, respectively, compared to doxorubicin’s IC50 values of 2.85 and 3.58 µM, respectively. The most active cytotoxic activity was exhibited by the target compounds (9b, 5a, 3a, 6b, 8b and 4a), listed in descending order according to their activity against the two cell lines, as represented in Figure 4. Interestingly, the attachment of the sulfanylmethyl-oxirane side chain at position-4 of the spiro cyclopenane-1,2′-pyridothienopyrimidine nucleus in compound 9b produced the most potent growth inhibition activity against both HepG2 and MCF-7 cancer cells, approximately 2–3 fold higher than doxorubicin, representing IC50 values of 1.17 and 1.52 μM, respectively. However, the activity slightly decreased against HepG2 cells and detectably decreased against MCF-7 for the spiro cyclohexane sulfanylmethyl oxirane analogue 9a compared to doxorubicin, with IC50 values of 4.88 and 23.56 µM, respectively. Furthermore, the incorporation of the morpholine nucleus into the pyridothienopyrimidine scaffold at position-2 in compound 5a led to a significant cytotoxic potency against both HepG2 and MCF-7, greater than that obtained from the reference drug, at IC50 values of 1.99 and 2.79 µM, respectively. However, the 4-methylpiperazinyl analogue 5b exhibited an observable reduction in growth inhibition activity towards both tested cancer cell lines with IC50 values of 10.16 and 21.06 µM, respectively. A comparable growth inhibitory activity to doxorubicin was displayed against HepG2 cancer cells by the 2,4-dichloro-pyridothienopyrimidine derivative 3a at an IC50 value of 2.31 μM, but its activity was less against MCF-7 with an IC50 value of 7.24 µM, while its 7-phenyl-9-p-tolyl analogue 3b represented a detectable drop in potency, with IC50 values of 11.34 and 24.72 µM against HepG2 and MCF-7, respectively. The spirocyclopentane-pyridothienopyrimidin-4-one derivative 6b was nearly equipotent to doxorubicin in inhibiting the growth of HepG2 cancer cells, with an IC50 value of 2.75 µM, but produced a mild decrease in activity against MCF-7 with IC50 of 9.89 µM; however, the displacement of the spiro pentane moiety with spirocyclohexane in the analogue 6d led to a twofold decrease in cytotoxic activity against HepG2 cells and a significant drop in potency against MCF-7 cancer cells with IC50 values of 4.45 and 21.67 µM, respectively. The other members in this series, 6a,c, appeared to be significantly less potent candidates than the reference drug, with IC50 ranges of 12.11–77.41 µM.
Additionally, the spiro cyclopentane sulfanylacetamide 8b produced an equivalent cytotoxic potency to that obtained by doxorubicin against HepG2 cells, with an IC50 value of 2.79 µM, but less potency than doxorubicin against MCF7, with an IC50 value of 13.54 µM. The spiro cyclohexane sulfanylacetamide analogue 8a appeared to be a less potent growth inhibitor towards both HepG2 and MCF-7 cancer cells, with IC50 values of 6.78 and 20.88 µM, respectively. Finally, the 2-chloromethyl-7,9-dimethyl derivative 4a, the last among the most potent compounds, showed significant growth inhibitory activity against HepG2 cancer cells with IC50 = 2.99 µM, but its potency decreased against MCF-7 cells with an IC50 value of 15.42 µM, whereas a high drop in activity was shown by the 7-phenyl-9-p-tolyl analogue 4b towards the two cancer cell lines with IC50 values of 36.52 and 43.27 µM, respectively. The rest of the target compounds, the spiro[cycloalkane-1,2′-pyridothienopyrimidine-4′(3′H)-thiones 7a–d and the pyridothienopyrimidine-2,4(1H,3H)-diones 2a,b, showed moderate to weak cytotoxic activities in the IC50 range of 10.35–41.62 µM, compared to doxorubicin.
The frequency and severity of undesirable effects on normal cells at therapeutic doses are considered among the most important characteristics that differentiate anticancer drugs from each other. Accordingly, the cytotoxic activity of the most active compounds (3a, 4a, 5a, 6b, 8b, 9b) was evaluated against the normal WISH cell line (Human amnion normal Liver cells) via MTT assay as IC50 values in μM, listed in Table 3. The obtained results revealed that the tested compounds had IC50 values in the range 394.98–460.23 µM, nearly equal to that obtained by the reference drug doxorubicin, IC50 doxorubicin 432.10 µM, which confirmed the high safety of the most potent compounds towards normal cells.

2.2.3. In Vitro EGFR Enzyme Inhibition Assay

The most potent cytotoxic compounds (3a, 4a, 5a, 6b, 8b, 9b) were subjected to further investigations of their inhibiting profiles against EGFR, in order to study one of their proposed modes of action as potent cytotoxic agents [54]. Accordingly, these derivatives were assessed as EGFR kinase inhibitors, using erlotinib as a reference drug as it is one of the most potent EGFR inhibitors [55]. The obtained results were expressed as IC50 values (nM) (Table 4). Interestingly, the most cytotoxic derivatives, spiro cyclopentane-1,2′ pyridothienopyrimidine–oxirane 9b and pyridothienopyrimidine–morpholine 5a, appeared to be 3–2 times more potent than erlotinib, representing IC50 values of 7.27, 9.66 and 27.01 nM, respectively. The 2-chloromethyl-pyridothienopyrimidin-4-one derivative 4a displayed a slight decrease in inhibition activity, but remained 1.5-fold more potent than erlotinib. An evident drop in EGFR inhibition activity was observed for the rest of the examined derivatives (3a, 6b, 8b) representing an IC50 range of 38.44–53.57 nM.

2.3. Molecular Docking Studies

Molecular docking studies were performed to study the binding modes of the most potent cytotoxic compounds (3a, 4a, 5a, 6b, 8b, 9b) to the active sites of the target EGFR compared with erlotinib (ERL) as EGFR inhibitor. Docking setup was first validated through self-docking of the co-crystallized ligand (ERL) in the vicinity of the binding site of the enzyme. The docking score (S) was –10.48 kcal/mol. and root mean square deviation (RMSD) was 1.03 Å (Figure 5). The calculated RMSD value confirms the validity of the docking procedure.
Examination of the binding interactions of erlotinib to the active site of the EGFR kinase domain showed several conventional hydrogen bond interactions with the Met769, Leu768, Val702 and Leu820 amino acids (Figure 6).
From the docking results of the examined compounds (Table 5), it was noticed that all compounds showed binding interactions within the active site of EGFR kinase domain with binding scores ranging from –12.01 to –8.94 kcal/mol. The spiro cyclopentane 4-((oxiran-2-ylmethyl)sulfanyl derivative 9b, which showed the highest biological activity, also showed the highest binding score of –12.01 kcal/mol, through binding with the key amino acids, Met769, Leu768, and Leu820, via the S atom of thiophene ring using a hydrogen bond acceptor, with further interactions with Thr766 via σ-hole bonding with the S atom of the S-methyl side chain. In addition, it bound to the amino acid Val702 through a hydrogen bond acceptor with the O atom of the oxirane moiety and showed high fitting in the vicinity of the active site, contributing to its high biological activity (Figure 7).
2-morpholinomethyl derivative 5a and 2,4-dichloro derivative 3a also showed higher binding scores than the co-crystallized ligand, erlotinib, at –11.48 and –11.42 kcal/mol, respectively; they showed a good binding pattern with the key amino acids, and compound 5a showed an ionic interaction with Asp831 revealing its high biological activity (Figure 8 and Figure 9). Other tested derivatives showed a lower binding score than erlotinib, with less binding interaction with the key amino acids.

3. Materials and Methods

3.1. Chemistry

3.1.1. General Information

The melting points were obtained in open capillary tubes using an Electrothermal IA9100 digital melting point apparatus. Elemental microanalyses were carried out at the Micro Analytical Unit at Cairo University. 1H NMR and 13C NMR spectra were recorded on a Bruker High Performance Digital FT-NMR Spectrometer Advance III (400/100 MHz) in the presence of TMS as the internal standard at Ain Shams University, Cairo, Egypt. Infrared spectra were measured using the KBr disc technique on a Jasco FT/IR-6100 Fourier transform IR spectrometer (Japan) at the scale of 400–4000 cm−1. ESI-mass spectra were determined using an Advion Compact Mass Spectrometer (CMS), NY, USA. TLC on silica gel-precoated aluminum sheets (Type 60, F 254, Merck, Darmstadt, Germany) was used for following up the reactions and checking the purity of the chemical compounds using chloroform/methanol (3:1, v/v), and spots were detected with iodine vapor or through exposure to a UV lamp at δ 254 nm for several seconds. The nomenclature of the compounds is according to the IUPAC system. The starting compounds, 3-amino-thieno[2,3-b]pyridine-2-carboxamides (1a,b), were prepared using the reported method [56].

3.1.2. Synthesis of Pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine-2,4(1H,3H)-diones 2a,b

A mixture of compounds 1a,b (5 mmol) and urea (0.42 g, 7 mmol) was heated at 180°C for 1 h. The solidified residue was treated with hot water and the obtained solid was separated by filtration. The solid was washed with hot water several times and recrystallized from acetone to yield compounds 2a,b.
7,9-Dimethylpyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine-2,4(1H,3H)-dione (2a) was obtained from 1a (1.11 g, 5 mmol) in 85% yield (1.05 g) as a brown solid, m.p. 325 °C. IR (KBr, υmax/cm−1): 3394, 3112 (2NH), 3010 (CH-aromatic), 2828 (CH-aliphatic), 1728, 1662 (2C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 2.55 (s, 3H, CH3), 2.79 (s, 3H, CH3), 7.21 (s, 1H, Ar-H), 10.59, 11.65 (2s, 2H, 2NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 20.40 (CH3), 24.36 (CH3), 108.64, 120.72, 122.85, 140.22, 145.63, 152.01, 160.07, 160.41, 161.48 (Ar-C, 2 C=O); ESI-MS: m/z = 247.33 [M − H+]. Anal. Calcd. for C11H9N3O2S (247.27): C, 53.43; H, 3.67; N, 16.99; S, 12.97% Found: C, 53.12, H, 3.91; N, 16.62; S, 13.31%.
7-Phenyl-9-(p-tolyl)pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine-2,4(1H,3H)-dione (2b) was obtained from 1b (1.79 g, 5 mmol) in 83% yield (1.60 g) as an orange solid, m.p. 316–317 °C. IR (KBr, υmax/cm−1): 3370, 3154 (2NH), 3024 (CH-aromatic), 2925, 2855 (CH-aliphatic), 1702, 1640 (2C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 2.43 (s, 3H, CH3), 7.46–7.52 (m, 5H, Ar-H), 7.60 (d, 2H, J = 8 Hz, Ar-H), 7.96 (s, 1H, Ar-H), 8.24 (d, 2H, J = 5.2 Hz, Ar-H), 10.42, 11.71 (2s, 2H, 2NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.42 (CH3), 112.45, 115.62, 118.27, 121.46, 127.21, 127.87, 129.30, 129.69, 130.23, 133.99, 138.39, 139.64, 144.46, 149.89, 151.96, 156.07, 157.40, 160.46 (Ar-C, 2 C=O); ESI-MS: m/z = 384.39 [M − H+]. Anal. Calcd. for C22H15N3O2S (385.44): C, 68.56; H, 3.92; N, 10.90; S, 8.32% Found: C, 68.74, H, 4.21; N, 10.68; S, 8.59%.

3.1.3. Synthesis of 2,4-Dichloropyrido[3′,2′:4,5]thieno[3,2-d]pyrimidines 3a,b

To a solution of compounds 2a,b (1 mmol) in phosphorus oxychloride (20 mL), phosphorus pentachloride (0.41 g, 2 mmol) was added. The reaction mixture was refluxed for 15 h, then left to cool. The mixture was poured slowly onto crushed ice and the formed solid was separated by filtration, washed with water several times, and recrystallized from ethanol to yield the 2,4-dichloro compounds 3a,b.
2,4-Dichloro-7,9-dimethylpyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine (3a) was obtained from 2a (0.25 g, 1 mmol) in 67% (0.19 g) yield as a buff solid, m.p. 250–251 °C. IR (KBr, υmax/cm−1): 3015 (CH-aromatic), 2920, 2853 (CH-aliphatic), 825, 768 (C-Cl); 1H-NMR (DMSO-d6, 400 MHz): δ = 2.59 (s, 3H, CH3), 2.76 (s, 3H, CH3), 7.34 (s, 1H, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 20.43 (CH3), 24.29 (CH3), 116.33, 120.63, 123.01, 140.22, 145.69, 152.10, 159.19, 161.22, 163.17 (Ar-C); ESI-MS: m/z = 283.13 [M − H+]. Anal. Calcd. for C11H7Cl2N3S (284.16): C, 46.50; H, 2.48; N, 14.79; S, 11.28% Found: C, 46.32, H, 2.59; N, 14.63; S, 11.09%.
2,4-Dichloro-7-phenyl-9-(p-tolyl)pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine (3b) was obtained from 2b (0.39 g, 1 mmol) in 73% yield (0.31 g) as a brown solid, m.p. 220 °C. IR (KBr, υmax/cm−1): 3094 (CH-aromatic), 2922 (CH-aliphatic), 1621 (C=N), 819, 768 (C-Cl); 1H-NMR (DMSO-d6, 400 MHz): δ = 2.47 (s, 3H, CH3), 7.34 (d, 2H, J = 8.4 Hz, Ar-H), 7.49–7.61 (m, 3H, Ar-H), 7.77 (d, 2H, J = 8.4 Hz, Ar-H), 8.14 (s, 1H, Ar-H), 8.24 (d, 2H, J = 3.6 Hz, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.38 (CH3), 113.52, 118.73, 119.18, 127.79, 129.21, 129.46, 130.38, 132.69, 133.23, 137.29, 138.95, 140.27, 150.70, 155.97, 157.27, 160.01, 161.26 (Ar-C); ESI-MS: m/z = 421.31 [M − H+]. Anal. Calcd. for C22H13Cl2N3S (422.33): C, 62.57; H, 3.10; N, 9.95; S, 7.59% Found: C, 62.76, H, 3.35; N, 9.73; S, 7.78 %.

3.1.4. Synthesis of 2-(Chloromethyl)pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4(3H)-ones 4a,b

To a cold solution of 1a,b (5 mmol) in acetonitrile (30 mL) at 0–5 °C, 2-chloroacetyl chloride (0.56 g, 5 mmol) was added dropwise while stirring. After addition, the reaction mixture was stirred for 1 h at room temperature and the solution was evaporated until dryness under reduced pressure, and then the oily residue was treated with hot petroleum ether 40–60. Then the formed solid was collected by filtration and recrystallized from ethanol to yield 4a,b.
2-(Chloromethyl)-7,9-dimethylpyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4(3H)-one (4a) was obtained from 1a (1.11 g, 5 mmol) in 78% yield (1.09 g) as a yellow solid, m.p. 330–331 °C. IR (KBr, υmax/cm−1): 3387 (NH), 3097 (CH-aromatic), 2821 (CH-aliphatic), 1669 (C=O), 772 (C-Cl); 1H-NMR (DMSO-d6, 400 MHz): δ = 2.59 (s, 3H, CH3), 2.88 (s, 3H, CH3), 4.66 (s, 2H, CH2Cl), 7.28 (s, 1H, Ar-H), 13.20 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 19.17, 24.37 (2CH3), 52.80 (CH2-Cl), 123.27, 124.88, 144.56, 147.42, 152.87, 158.65, 159.42, 160.22 (Ar-C), 161.93 (C=O); ESI-MS: m/z = 278.69 [M − H+]. Anal. Calcd. for C12H10ClN3OS (279.74): C, 51.52; H, 3.60; N, 15.02; S, 11.46% Found: C, 51.28; H, 3.35; N, 14.82; S, 11.20%.
2-(Chloromethyl)-7-phenyl-9-(p-tolyl)pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4(3H)-one (4b) was obtained from 1b (1.79 g, 5 mmol) in 82% yield (1.71 g) as a yellow solid, m.p. 310 °C. IR (KBr, υmax/cm−1): 3437 (NH), 3119 (CH-aromatic), 2920 (CH-aliphatic), 1669 (C=O), 776 (C-Cl); 1H-NMR (DMSO-d6, 400 MHz): δ = 2.43 (s, 3H, CH3), 4.38 (s, 2H, CH2Cl), 7.32 (d, 2H, J = 6.4 Hz, Ar-H), 7.45–7.57 (m, 3H, Ar-H), 7.66 (d, 2H, J = 6.4 Hz, Ar-H), 7.99 (s, 1H, Ar-H), 8.21 (d, 2H, J = 8.4 Hz, Ar-H), 9.79 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.32 (CH3), 53.71 (CH2-Cl), 119.37, 126.63, 127.15, 127.94, 129.16, 129.51, 130.08, 132.64, 139.94, 140.57, 147.84, 148.12, 149.61,154.64, 155.69, 159.02, 160.33 (Ar-C, C=O); ESI-MS: m/z = 416.87 [M − H+]. Anal. Calcd. for C23H16ClN3OS (417.91): C, 66.10; H, 3.86; N, 10.06; S, 7.67% Found: C, 66.41, H, 4.06; N, 9.86; S, 7.88%.

3.1.5. Synthesis of 2-Substituted-7-phenyl-9-(p-tolyl)pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4(3H)-ones 5a,b

A mixture of 4b (0.42 g, 1 mmol) and the appropriate amine (1 mmol) in DMF (15 mL) was refluxed for 1 hr, and then the reaction mixture was poured onto cold water. The obtained precipitate was separated by filtration, washed with water, and recrystallized from ethanol to yield 5a,b.
2-(Morpholinomethyl)-7-phenyl-9-(p-tolyl)pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4(3H)-one (5a) was obtained by reaction of 4b with morpholine (0.087 g, 1mmol) in 79% yield (0.37 g) as a pale yellow solid, m.p. 245 °C. IR (KBr, υmax/cm−1): 3440 (NH), 3094 (CH-aromatic), 2919, 2854 (CH-aliphatic), 1662 (C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 2.38 (s, 4H, 2CH2N-morpholine), 2.39 (s, 3H, CH3), 3.29 (s, 2H, CH2N), 3.54 (s, 4H, 2CH2O), 7.28 (d, 2H, J = 10.4 Hz, Ar-H), 7.49–7.60 (m, 3H, Ar-H), 7.63 (d, 2H, J =10.4 Hz, Ar-H), 7.97 (s, 1H, Ar-H), 8.27 (d, 2H, J = 8.4 Hz, Ar-H), 12.66 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.44 (CH3), 53.37 (2CH2N-morpholine), 60.65 (CH2N), 66.56 (2CH2O), 119.80, 127.49, 127.82, 128.47, 129.49, 130.45, 130.83, 132.14, 139.71, 140.50, 150.02, 151.48, 155.65, 157.06, 158.75 (Ar-C), 163.50 (C=O); ESI-MS: m/z = 467.50 [M − H+]. Anal. Calcd. for C27H24N4O2S (468.58): C, 69.21; H, 5.16; N, 11.96; S, 6.84 % Found: C, 69.48; H, 5.39; N, 11.72; S, 6.51%.
2-((4-Methylpiperazin-1-yl)methyl)-7-phenyl-9-(p-tolyl)pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4(3H)-one (5b) was obtained by reaction of 4b with 4-methylpiperazine (0.100 g, 1mmol) in 73% yield (0.35 g) as a yellowish white solid, m.p. 260–261 °C. IR (KBr, υmax/cm−1): 3433 (NH), 3032 (CH-aromatic), 2920 (CH-aliphatic), 1655 (C=O); 1H-NMR (CDCl3, 400 MHz): δ = 2.48 (s, 3H, CH3), 2.54 (s, 3H, NCH3), 2.81–3.01 (m, 8H, 4CH2N piperazine), 3.57 (s, 2H, NCH2), 7.32 (d, 2H, J = 8 Hz, Ar-H), 7.48–7.54 (m, 3H, Ar-H), 7.64 (d, 2H, J = 8.4 Hz, Ar-H), 7.81 (s, 1H, Ar-H), 8.17 (d, 2H, J = 6.8 Hz, Ar-H), 11.32 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.34 (CH3), 48.22 (CH3N), 56.13, 57.56 (4CH2N-piperazine), 61.49 (CH2N), 119.11, 126.96, 127.21, 127.79, 128.90, 129.01, 129.44, 130.08, 132.56, 139.95, 140.21, 145.01, 149.47, 150.17, 155.56, 157.15, 158.44 (Ar-C), 163.25 (C=O); ESI-MS: m/z = 480.64 [M − H+]. Anal. Calcd. for C28H27N5OS (481.62): C, 69.83; H, 5.65; N, 14.54; S, 6.66% Found: C, 69.62, H, 5.90; N, 14.21; S, 6.79%.

3.1.6. Synthesis of 1′H-Spiro[cycloalkane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′(3′H)-ones 6a–d

A mixture of 1a,b (0.02 mol)) and the appropriate cyclic ketone (0.03 mol) in DMF (50 mL) containing zinc chloride anhydrous (2.72 g, 0.02 mol) was heated under reflux for 6 h, and then the reaction mixture was poured onto cold water. The obtained precipitate was separated by filtration, washed with water, and recrystallized from DMF/H2O to yield 6a–d.
7′,9′-Dimethyl-1′H-spiro[cyclopentane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′(3′H)-one (6a) was obtained by reaction of 1a (4.42 g, 0.02 mol) with cyclopentanone (2.52 g, 0.03 mol) in 75% yield (4.31 g) as a brown solid, m.p. 245–246 °C. IR (KBr, υmax/cm−1): 3319, 3210 (2NH), 3070 (CH-aromatic), 2921, 2828 (CH-aliphatic), 1650 (C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.72–2.07 (m, 8H, 4CH2), 2.51 (s, 3H, CH3), 2.66)s, 3H, CH3), 6.48 (s, 1H, NH, D2O exchangeable), 7.04 (s,1H, Ar-H), 9.82 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 20.41 (CH3), 22.65, (2CH2), 24.33 (CH3), 37.55 (2CH2), 70.15 (spiro C), 122.54, 123.29, 135.88, 145.63, 148.71, 159.30, 161.57 (Ar-C), 167.65 (C=O); ESI-MS: m/z = 286.41 [M − H+]. Anal. Calcd. for C15H17N3OS (287.38): C, 62.69; H, 5.96; N, 14.62; S, 11.16% Found: C, 62.52, H, 5.74; N, 14.41; S, 10.98%.
7′-Phenyl-9′-(p-tolyl)-1′H-spiro[cyclopentane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′(3′H)-one (6b) was obtained by reaction of 1b (7.19 g, 0.02 mol) with cyclopentanone (2.52 g, 0.03 mol) in 71% yield (6.04 g) as a yellow solid, m.p. 210 °C. IR (KBr, υmax/cm−1): 3410, 3200 (2NH), 3055 (CH-aromatic), 2919 (CH-aliphatic), 1644 (C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.21–1.85 (m, 8H, 4CH2), 2.43 (s, 3H, CH3), 5.94 (s, 1H, NH, D2O exchangeable), 7.29 (d, 2H, J = 7.4 Hz, Ar-H), 7.41–7.62 (m, 5H, Ar-H), 7.81(s, 1H, NH, D2O exchangeable), 7.96 (s, 1H, Ar-H), 8.22 (d, 2H, J = 10.2 Hz, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.36, 21.54 (2CH2, CH3), 38.33 (2CH2), 70.81 (spiro C), 118.26, 121.25, 127.54, 127.65, 129.39, 129.79, 130.32, 133.82, 139.15, 139.31, 148.17, 150.62, 155.15, 163.25 (Ar-C), 167.38 (C=O); ESI-MS: m/z = 424.50 [M − H+]. Anal. Calcd. for C26H23N3OS (425.55): C, 73.38; H, 5.45; N, 9.87; S, 7.53% Found: C, 73.64, H, 5.21; N, 10.08; S, 7.38%.
7′,9′-Dimethyl-1′H-spiro[cyclohexane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′(3′H)-one (6c) was obtained by reaction of 1a (4.42 g, 0.02 mol) with cyclohexanone (2.94 g, 0.03 mol) in 79% yield (4.76 g) as a brown solid, m.p. 290–291 °C. IR (KBr, υmax/cm−1): 3317, 3168 (2NH), 3043 (CH-aromatic), 2932 (CH-aliphatic), 1649 (C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.23–2.06 (m, 10H, 5CH2), 2.51 (s, 3H, CH3), 2.71 (s, 3H, CH3), 5.91 (s, 1H, NH, D2O exchangeable), 7. 07 (s, 1H, Ar-H), 7.86 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 19.36, 20.16, 22.27, 24.32 (3CH2, 2CH3), 36.24 (2CH2), 69.49 (spiro C), 122.11, 123.58, 134.96, 145.31, 148.44, 159.14, 161.51 (Ar-C), 167.80 (C=O); ESI-MS: m/z = 300.46 [M − H+]. Anal. Calcd. for C16H19N3OS (301.41): C, 63.76; H, 6.35; N, 13.94; S, 10.64% Found: C, 63.49, H, 6.17; N, 13.69; S, 10.41%.
7′-Phenyl-9′-(p-tolyl)-1′H-spiro[cyclohexane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′(3′H)-one (6d) was obtained by reaction of 1b (7.19 g, 0.02 mol) with cyclohexanone (2.94 g, 0.03 mol) in 77% yield (6.77 g) as a yellow solid, m.p. 263 °C. IR (KBr, υmax/cm−1): 3415, 3166 (2NH), 3053 (CH-aromatic), 2924, 2854 (CH-aliphatic), 1660 (C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.34–1.87 (m, 10H, 5CH2), 2.42 (s, 3H, CH3), 4.59 (s, 1H, NH, D2O exchangeable), 7.38 (d, 2H, J = 6.8 Hz, Ar-H), 7.49–7.54 (m, 5H, Ar-H), 7.83 (s, 1H, Ar-H), 8.06 (s, 1H, NH, D2O exchangeable), 8.21 (d, 2H, J = 6 Hz, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 19.96, 21.38, 22.54 (3CH2, CH3), 35.86 (2CH2), 70.67 (spiro C), 118.52, 121.27, 127.48, 127.84, 129.24, 129.67, 129.96, 130.86, 133.48, 137.29, 139.20, 139.50, 148.67, 155.50, 156.14 (Ar-C), 167.73 (C=O); ESI-MS: m/z = 438.52 [M − H+]. Anal. Calcd. for C27H25N3OS (439.58): C, 73.77; H, 5.73; N, 9.56; S, 7.29% Found: C, 73.54, H, 5.49; N, 9.69; S, 7.59%.

3.1.7. Synthesis of 1′H-Spiro[cycloalkane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine]-4′(3′H)-thiones 7a–d

A mixture of 6a–d (5 mmol) and phosphorus pentasulfide (1.11 g, 5 mmol) in pyridine (40 mL) was heated under reflux for 12 h. The reaction mixture was poured onto cold water and left overnight. The formed solid was separated by filtration, washed with water, and recrystallized from 1,4-dioxane to yield compounds 7a–d.
7′,9′-Dimethyl-1′H-spiro[cyclopentane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine]-4′(3′H)-thione (7a) was obtained from 6a (1.44 g, 5 mmol) in 63% yield (0.95 g) as a burnt orange solid, m.p. 220–221 °C. IR (KBr, υmax/cm−1): 3372, 3157 (2NH), 3063 (CH-aromatic), 2922 (CH-aliphatic), 1247 (C=S); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.36–2.03 (m, 8H, 4CH2), 2.57 (s, 3H, CH3), 2.73 (s, 3H, CH3), 6.49 (s, 1H, NH,D2O exchangeable), 7.03 (s,1H, Ar-H), 9.85 (s, 1H, NH,D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ =19.56, 22.53 (CH3, 2CH2), 24.43 (CH3), 37.67 (2CH2), 78.59 (spiro C), 117.93, 122.11, 123.15, 139.87, 146.17, 159.21, 162.86 (Ar-C),182.25 (C=S); ESI-MS: m/z = 302.40 [M − H+]. Anal. Calcd. for C15H17N3S2 (303.44): C, 59.37; H, 5.65; N, 13.85; S, 21.13% Found: C, 59.09, H, 5.47; N, 13.56; S, 20.89%.
7′-Phenyl-9′-(p-tolyl)-1′H-spiro[cyclopentane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine]-4′(3′H)-thione (7b) was obtained from 6b (2.13 g, 5 mmol) in 68% yield (1.50 g) as a brown solid, m.p. 180 °C. IR (KBr, υmax/cm−1): 3402, 3186 (2NH), 3055 (CH-aromatic), 2923, 2852 (CH-aliphatic), 1232 (C=S); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.34–2.01 (m, 8H, 4CH2), 2.42 (s, 3H, CH3), 4.67 (s, 1H, NH, D2O exchangeable), 7.42 (d, 2H, J = 7.2 Hz, Ar-H), 7.48–7.53 (m, 5H, Ar-H), 7.83 (s, 1H, Ar-H), 8.22 (d, 2H, J = 7.2 Hz, Ar-H), 9.68 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.47, 21.96 (2CH2, CH3), 36.50 (2CH2), 77.79 (spiro C), 118.42, 121.20, 124.67, 127.47, 127.71, 129.11, 129.74, 130.20, 133.56, 139.21, 139.34, 148.17, 149.62, 155.15, 163.25 (Ar-C), 183.35 (C=S); ESI-MS: m/z = 440.69 [M − H+]. Anal. Calcd. for C26H23N3S2 (441.61): C, 70.72; H, 5.25; N, 9.52; S, 14.52% Found: C, 70.53, H, 5.01; N, 9.37; S, 14.38%.
7′,9′-Dimethyl-1′H-spiro[cyclohexane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine]-4′(3′H)-thione (7c) was obtained from 6c (1.51 g, 5 mmol) in 71% yield (1.13 g) as a greenish grey solid, m.p. 201 °C. IR (KBr, υmax/cm−1): 3427, 3190 (2NH), 3048 (CH-aromatic), 2929, 2853 (CH-aliphatic), 1234 (C=S); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.21–2.11 (m, 10H, 5CH2), 2.51 (s, 3H, CH3), 2.71 (s, 3H, CH3), 6.22 (s, 1H, NH, D2O exchangeable), 7. 07 (s, 1H, Ar-H), 9.68 (s, 1H, NH, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 19.26, 22.02, 24.32, 25.15 (3CH2, 2CH3), 34.46 (2CH2), 70.16 (spiro C), 118.38,122.29, 123.61, 138.66, 146.42, 159.31, 162.71 (Ar-C), 181.33 (C=S); ESI-MS: m/z = 316.40 [M − H+]. Anal. Calcd. for C16H19N3S2 (317.47): C, 60.53; H, 6.03; N, 13.24; S, 20.20% Found: C, 60.79, H, 6.17; N, 13.49; S, 20.44%.
7′-Phenyl-9′-(p-tolyl)-1′H-spiro[cyclohexane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine]-4′(3′H)-thione (7d) was obtained from 6d (2.19 g, 5 mmol) in 66% yield (1.50 g) as a yellow solid, m.p. 178–179 °C. IR (KBr, υmax/cm−1): 3402, 3188 (NH), 3068 (CH-aromatic), 2923, 2852 (CH-aliphatic), 1232 (C=S); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.2 4–2.07 (m, 10H, 5CH2), 2.43 (s, 3H, CH3), 4.79 (s, 1H, NH, D2O exchangeable), 7.39 (d, 2H, J = 8.4 Hz, Ar-H), 7.48–7.57 (m, 5H, Ar-H), 7.87 (s, 1H, Ar-H), 9.66 (s, 1H, NH, D2O exchangeable), 8.22 (d, 2H, J = 6.4 Hz, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.22, 21.47, 24.46 (3CH2, CH3), 34.87 (2CH2), 70.46 (spiro C), 118.21, 121.45, 124.25, 127.39, 127.73, 129.39, 129.66, 130.42, 133.92, 137.90, 139.96, 140.00, 148.82, 156.27, 163.83 (Ar-C), 182.12 (C=S); ESI-MS: m/z = 454.60 [M − H+]. Anal. Calcd. for C27H25N3S2 (455.64): C, 71.17; H, 5.53; N, 9.22; S, 14.07% Found: C, 71.36, H, 5.19; N, 9.46; S, 13.89%.

3.1.8. Synthesis of 2-((1′H-Spiro[cycloalkane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′-yl)sulfanyl)acetamides 8a,b

A mixture of compounds 7c,b (2 mmol) and 2-chloroacetamide (0.187 g, 2 mmol) in DMF (30 mL) containing sodium carbonate anhydrous (0.5 g) was refluxed for 6 h. The reaction mixture was poured onto iced water and the medium was neutralized with 1N HCl to pH = 7. The formed precipitate was separated by filtration, washed with water, and recrystallized from chloroform to yield the acetamide derivatives 8a,b.
2-((7′,9′-Dimethyl-1′H-spiro[cyclohexane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′-yl)sulfanyl)acetamide (8a) was obtained from 7c (0.63 g, 2 mmol) in 73% yield (0.55 g) as a brown solid, m.p. 160–161 °C. IR (KBr, υmax/cm−1): 3403, 3233 (NH), 3048 (CH-aromatic), 2926, 2854 (CH-aliphatic), 1674 (C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.22–2.09 (m, 10H, 5CH2), 2.68 (s, 3H, CH3), 2.76 (s, 3H, CH3), 4.05 (s, 2H, SCH2), 6.24 (s, 1H, NH, D2O exchangeable), 7. 09 (s, 1H, Ar-H), 7.35 (s, 2H, NH2, D2O exchangeable); 13C-NMR (DMSO-d6, 100 MHz): δ = 19.21, 22.71, 24.73, 25.41 (3CH2, 2CH3), 36.82 (2CH2), 41.93 (SCH2), 72.43 (spiro C), 118.78, 122.54, 123.01, 124.67, 144.63, 147.53, 159.35, 163.15 (Ar-C), 168.23 (C=O); ESI-MS: m/z = 373.50 [M − H+]. Anal. Calcd. for C18H22N4OS2 (374.52): C, 57.73; H, 5.92; N, 14.96; S, 17.12% Found: C, 57.49, H, 5.18; N, 14.58; S, 17.43%.
2-((7′-Phenyl-9′-(p-tolyl)-1′H-spiro[cyclopentane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin]-4′-yl)sulfanyl)acetamide (8b) was obtained from 7b (0.88 g, 2 mmol) in 69% yield (0.69 g) as a brown solid, m.p. 148–149 °C. IR (KBr, υmax/cm−1): 3400, 3200 (NH), 3054 (CH-aromatic), 2918, 2853 (CH-aliphatic), 1667 (C=O); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.21–2.03 (m, 8H, 4CH2), 2.43 (s, 3H, CH3), 4.19 (s, 2H, SCH2), 6.86 (s, 1H, NH, D2O exchangeable), 7.34 (d, 2H, J = 5.6 Hz, Ar-H), 7.44–7.63 (m, 5H, Ar-H), 7.72 (s, 2H, NH2, D2O exchangeable), 7.83 (s, 1H, Ar-H), 8.24 (d, 2H, J = 6.8 Hz, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.43 (CH3), 29.61(2CH2), 33.86 (2CH2), 41.41(SCH2), 74.27 (spiro C), 118.55, 122.05, 126.94, 127.85, 127.96, 128.90, 129.50, 130.44, 133.49, 139.05, 140.57, 147.93, 150.02,155.45, 157.40, 163.26 (Ar-C), 169.38 (C=O); ESI-MS: m/z = 497.68 [M − H+]. Anal. Calcd. for C28H26N4OS2 (498.66): C, 67.44; H, 5.26; N, 11.24; S, 12.86% Found: C, 67.63, H, 5.44; N, 11.50; S, 13.05%.

3.1.9. Synthesis of 4′-((oxiran-2-ylmethyl)sulfanyl)-1′H-spiro[cycloalkane-1,2′-pyrido[3′,2′:4,5]thieno-[3,2-d]pyrimidine] 9a,b

A mixture of compounds 7c,b (1mmol)) and epichlorohydrin (0.092 g, 1mmol) in acetone (20 mL) containing triethyl amine (0.1 mL) was refluxed for 5 h. The solvent was evaporated until dryness and the oily residue was treated with hot petroleum ether (30 mL). The formed solid was separated by filtration and recrystallized from ethanol to yield 9a,b.
7′,9′-Dimethyl-4′-((oxiran-2-ylmethyl)sulfanyl)-1′H-spiro[cyclohexane-1,2′-pyrido[3′,2′:4,5] thieno[3,2 -d]pyrimidine] (9a), was obtained from 7c (0.32 g, 1 mmol) in 78% yield (0.29 g) as a grey solid, m.p. 155 °C. IR (KBr, υmax/cm−1): 3430 (NH), 3040 (CH-aromatic), 2928, 2856 (CH-aliphatic), 1623 (C=N); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.24–1.89 (m, 10H, 5CH2), 2.60 (s, 3H, CH3), 2.74 (s, 3H, CH3), 2.86–2.94 (2m, 2H, OCH2- oxirane), 3.67 (d, 2H, J = 10.8 Hz, SCH2), 3.72–3.77 (m, 1H, OCH-oxirane), 6.12 (s, 1H, NH, D2O exchangeable), 7. 06 (s, 1H, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 19.27, 22.91, 24.70, 25.33 (3CH2, 2CH3), 32.19 (SCH2), 37.02 (2CH2), 44.76, 52.45 (OCH2, OCH, oxirane), 70.03 (spiro C), 118.74, 122.33, 123.12, 124.67, 145.00, 147.21, 159.19, 163.66 (Ar-C); ESI-MS: m/z = 372.60 [M − H+]. Anal. Calcd. for C19H23N3OS2 (373.53): C, 61.09; H, 6.21; N, 11.25; S, 17.17% Found: C, 61.23, H, 6.44; N, 11.51; S, 17.01%.
4′-((Oxiran-2-ylmethyl)sulfanyl)-7′-phenyl-9′-(p-tolyl)-1′H-spiro[cyclopentane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine] (9b) was obtained from 7b (0.44 g, 1 mmol) in 74% yield (0.37 g) as a pale yellow solid, m.p. 120 °C. IR (KBr, υmax/cm−1): 3422 (NH), 3094 (CH-aromatic), 2919 (CH-aliphatic), (C=N); 1H-NMR (DMSO-d6, 400 MHz): δ = 1.35–1.94 (m, 8H, 4CH2), 2.43 (s, 3H, CH3), 2.94–3.13 (2m, 2H, OCH2- oxirane), 3.51 (d, 2H, J = 6.0 Hz, SCH2), 3.91–3.95 (m, 1H, OCH-oxirane), 6.87 (s, 1H, NH, D2O exchangeable), 7.32 (d, 2H, J = 7.2 Hz, Ar-H), 7.48–7.69 (m, 5H, Ar-H), 7.87 (s, 1H, Ar-H), 8.26 (d, 2H, J = 7.6 Hz, Ar-H); 13C-NMR (DMSO-d6, 100 MHz): δ = 21.40 (CH3), 29.15 (2CH2), 31.88 (SCH2), 36.33 (2CH2), 46.33, 53.45 (OCH2, OCH, oxirane), 69.33 (spiro C), 118.49, 123.62, 127.81, 127.93, 128.78, 128.94, 129.43, 129.50, 130.53, 133.77, 139.17, 140.27, 148.73, 152.73,156.17, 163.83 (Ar-C); ESI-MS: m/z = 496.70 [M − H+]. Anal. Calcd. for C29H27N3OS2 (497.68): C, 69.99; H, 5.47; N, 8.44; S, 12.88% Found: C, 69.81, H, 5.31; N, 8.20; S, 12.63 %.

3.2. Antimicrobial Assay

All synthesized compounds (2a,b9a,b) were screened for their in vitro antimicrobial activity against five bacterial strains (S. aureus 25923, B. subtilis 6633, B. cereus 33018, E. coli 8739, S. typhimurium 14028), three yeasts (C. albicans 10231, C. tropicals 750, S. cerevisiae) and two fungi (A. flavus, A. niger EM77). The MIC values (in μg/mL) of the tested compounds were determined using the broth dilution method and are listed in Table 1 and Table 2 [57]. (More details are provided in the Supplementary Materials).

3.3. In Vitro Cytotoxicity Screening

The in vitro cytotoxic activity of the target compounds 2a,b9a,b was screened against HepG-2 and MCF-7 cancer cell lines, as well as the WISH normal cell line for the most active compounds, using the MTT assay [53]. The cells used in the cytotoxicity assays were cultured in a RPMI 1640 medium supplemented with 10% fetal calf serum. The cytotoxicity was estimated as IC50 in μM for the tested compounds and the reference drug doxorubicin, and are listed in Table 3. (More details are provided in Supplementary Materials).

3.4. EGFR Kinase Inhibitory Assay

EGFR kinase inhibitory assays were performed for target compounds 3a, 4a, 5a, 6b, 8b and 9b with erlotinib as a reference inhibitor using the EGFR kinase assay kit (Cat. # 40321). The assay kit is designed to measure EGFR Kinase activity for screening applications using Kinase-Glo® MAX as a detection reagent. The luminescence was measured using a microplate reader (Infinite M200 microplate reader, Tecan, Männedorf, Switzerland) [54]. All assays were performed in triplicate and the relative inhibition (%) of inhibitors was then calculated via comparison with the control with no inhibitor. Then, the IC50 values (the concentration that provides 50% enzyme inhibition) and their standard deviation (SD) for the tested compounds and the reference drug were determined in μM and are listed in Table 4. (More details are provided in the Supplementary Materials).

3.5. Molecular Modeling Studies

To investigate molecular interactions between the most potent compounds and the active site of the epidermal growth factor receptor (EGFR), molecular docking study was performed using molecular operating environment software (MOE 2019.0102). Energy minimization was carried out until a RMSD gradient of 0.1 kcal∙mol−1Å−1 was achieved using a MMFF94x force field. The co-crystalized ligand (Erlotinib) was used to define the binding site for docking [58,59]. (More details are provided in the Supplementary Materials).

4. Conclusions

In conclusion, two novel series of pyridothienopyrimidine and spiro[cyclopentane/hexane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine] derivatives 2a,b5a,b and 6a,b9a,b, were synthesized and structurally elucidated. The new derivatives were subjected to in vitro antimicrobial screening against a panel of bacterial and fungal pathogens. According to the MIC values, derivatives 3a, 4a,b, 5a, 6b,c, 7c, 8b, and 9b exhibited significant antibacterial and antifungal activity with MIC ranges of 4–16 μg/mL, compared to amoxicillin trihydrate and clotrimazole as reference drugs with MIC ranges of 4–16 µM.
Furthermore, all new derivatives were evaluated as cytotoxic agents against HepG2 and MCF7 cancer cell lines. Compounds 9b, 5a, 3a, 6b, 8b, 4a produced the most potent antiproliferative activity with IC50 values ranging between 1.17–2.99 µM against HepG2 cells and IC50s ranging between 1.52–15.42 µM against MCF-7 cells, compared to doxorubicin as reference drug with IC50s of 2.85 and 3.58 µM, respectively. In addition, these derivatives exhibited a promising safety profile when evaluated against the human normal WISH cell line.
Compounds 3a, 4a, 5a, 6b, 8b, 9b presented promising dual antimicrobial and antiproliferative activities, achieving the desired goal of this study.
The suppressing effect of compounds 3a, 4a, 5a, 6b, 8b, 9b against EGFR TK was also evaluated. It was detected that compounds 9b, 5a, 4a showed higher suppressing activity than erlotinib with IC50 values of 7.27, 9.66 and 27.01 nM, respectively. In addition, molecular docking study was performed to determine the modes of interaction of the examined derivatives with amino acid residues at the active site of EGFR-PK. The docking results revealed that compounds 9a and 5a showed higher binding scores (–12.01 and –11.48 kcal/mol) than that of erlotinib (–10.48 kcal/mol).

Supplementary Materials

Figures S2–S34: NMR spectral data of the new compounds, S35: in vitro antimicrobial assay, S35: cytotoxicity MTT assay, S36: in vitro EGFR kinase inhibitory assay, S37: docking modeling evaluation.

Author Contributions

Conceptualization, E.M.M.E.-D.; supervision, E.M.M.E.-D., M.M.A., R.R.K.; investigation, E.M.M.E.-D., A.A.A.E.-G., E.A.K.; methodology, A.A.A.E.-G., E.A.K.; software, M.K.E.-A.; data curation, E.M.M.E.-D., A.A.A.E.-G.; writing—original draft preparation, E.M.M.E.-D., A.A.A.E.-G., M.M.A., writing—review and editing, E.M.M.E.-D., A.A.A.E.-G., M.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in supplementary material.

Acknowledgments

Authors are grateful to National Research Centre for its support of this work.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds 5a, 6c and 7c are available from the authors.

References

  1. Hay, S.I.; Rao, P.C.; Dolecek, C.; Day, N.P.J.; Stergachis, A.; Lopez, A.D.; Murray, C.J.L. Measuring and mapping the global burden of antimicrobial resistance. BMC Med. 2018, 16, 78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Aslam, B.; Wang, W.; Arshad, M.I.; Khurshid, M.; Muzammil, S.; Rasool, M.H.; Nisar, M.A.; Alvi, R.F.; Aslam, M.A.; Qamar, M.U.; et al. Antibiotic resistance: A rundown of a global crisis. Infect. Drug Resist. 2018, 11, 1645–1658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Lomazzi, M.; Moore, M.; Johnson, A.; Balasegaram, M.; Borisch, B. Antimicrobial resistance—Moving forward? BMC Public Health 2019, 19, 858. [Google Scholar] [CrossRef] [PubMed]
  4. Varela, M.F.; Stephen, J.; Lekshmi, M.; Ojha, M.; Wenzel, N.; Sanford, L.M.; Hernandez, A.J.; Parvathi, A.; Kumar, S.H. Bacterial Resistance to Antimicrobial Agents. Antibiotics 2021, 10, 593. [Google Scholar] [CrossRef]
  5. Annunziato, G. Strategies to Overcome Antimicrobial Resistance (AMR) Making Use of Non-Essential Target Inhibitors: A Review. Int. J. Mol. Sci. 2019, 20, 5844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Liu, B.; Jiang, D.; Hu, G. The antibacterial activity of isatin hybrids. Curr. Top Med. Chem. 2021, 22, 25–40. [Google Scholar] [CrossRef]
  7. Wang, X.; Zhang, H.; Chen, X. Drug resistance and combating drug resistance in cancer. Cancer Drug Resist. 2019, 2, 141–160. [Google Scholar] [CrossRef] [Green Version]
  8. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  9. Presti, D.; Quaquarini, E. The PI3K/AKT/mTOR and CDK4/6 Pathways in Endocrine Resistant HR+/HER2- Metastatic Breast Cancer: Biological Mechanisms and New Treatments. Cancers 2019, 11, 1242. [Google Scholar] [CrossRef] [Green Version]
  10. Sharma, P.; Kaur, S.; Chadha, B.S.; Kaur, R.; Kaur, M.; Kaur, S. Anticancer and antimicrobial potential of enterocin 12a from Enterococcus faecium. BMC Microbiol. 2021, 21, 39. [Google Scholar] [CrossRef]
  11. Felício, M.R.; Silva, O.N.; Gonçalves, S.; Santos, N.C.; Franco, O.L. Peptides with dual antimicrobial and anticancer activities. Front. Chem. 2017, 5, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Hosseinzadeh, Z.; Ramazani, A.; Razzaghi-Asl, N. Anti-cancer nitrogen-containing heterocyclic compounds. Curr. Org. Chem. 2018, 22, 2256–2279. [Google Scholar] [CrossRef]
  13. Mermer, A.; Keles, T.; Sirin, Y. Recent Studies of Nitrogen Containing Heterocyclic Compounds as Novel Antiviral Agents: A Review. Bioorganic Chem. 2021, 114, 105076. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, S.; Yuan, X.H.; Wang, S.Q.; Zhao, W.; Chen, X.B.; Yua, B. FDA-approved pyrimidine-fused bicyclic heterocycles for cancer therapy: Synthesis and clinical application. Eur. J. Med. Chem. 2021, 2141, 13218. [Google Scholar] [CrossRef]
  15. Mohi El-Deen, E.M.; Anwar, M.M.; Hasabelnaby, S.M. Synthesis and in vitro cytotoxic evaluation of some novel hexahydroquinoline derivatives containing benzofuran moiety. Res. Chem. Intermed. 2016, 42, 1863–1883. [Google Scholar] [CrossRef]
  16. Bhutani, P.; Joshi, G.; Raja, N.; Bachhav, N.; Rajanna, P.K.; Bhutani, H.; Paul, A.T.; Kumar, R.U.S. FDA Approved Drugs from 2015–June 2020: A Perspective. J. Med. Chem. 2021, 64, 2339–2381. [Google Scholar] [CrossRef]
  17. Abdelaziz, M.E.; El-Miligy, M.M.M.; Fahmy, S.M.; Mahran, M.A.; Hazzaa, A.A. Design, synthesis and docking study of pyridine and thieno[2,3-b] pyridine derivatives anticancer PIM-1 kinase inhibitors. Bioorg. Chem. 2018, 80, 674–692. [Google Scholar] [CrossRef]
  18. Eurtivong, C.; Semenov, V.; Semenova, M.; Konyushkin, L.; Atamanenko, O.; Reynisson, J.; Kiselyov, A. 3-Amino-thieno[2,3-b]pyridines as microtubule-destabilising agents: Molecular modelling and biological evaluation in the sea urchin embryo and human cancer cells. Bioorg. Med. Chem. 2017, 25, 658–664. [Google Scholar] [CrossRef]
  19. Al-Trawneh, S.A.; Tarawneh, A.H.; Gadetskaya, A.V.; Seo, E.; Al-Ta’ani, M.R.; Al-Taweel, S.A.; El-Abadelah, M.M. Synthesis and cytotoxicity of thieno[2,3-b]pyridine derivatives toward sensitive and multidrug-resistant leukemia cells. Acta Chim. Solv. 2021, 68, 458–465. [Google Scholar] [CrossRef]
  20. Elsherif, M.A. Antibacterial evaluation and molecular properties of pyrazolo[3,4-b]pyridines and thieno[2,3-b] pyridines. J. Appl. Pharm. Sci. 2021, 11, 118–124. [Google Scholar] [CrossRef]
  21. Mohi El-Deen, E.M.; Abd El-Meguid, E.A.; Hasabelnaby, S.; Karam, E.A.; Nossier, E.S. Synthesis, Docking Studies, and In Vitro Evaluation of Some Novel Thienopyridines and Fused Thienopyridine–Quinolines as Antibacterial Agents and DNA Gyrase Inhibitors. Molecules 2019, 24, 3650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Mekky, A.E.M.; Sanad, S.M.H.; Said, A.Y.; Elneairy, M.A.A. Synthesis, cytotoxicity, in-vitro antibacterial screening and in-silico study of novel thieno[2,3-b]pyridines as potential pim-1 inhibitors. Synth. Commun. 2020, 50, 2376–2389. [Google Scholar] [CrossRef]
  23. Attaby, F.A.; Abdel-Fattah, A.M.; Shaif, L.M.; Elsayed, M.M. Reactions, Anti-Alzheimer and Anti COX-2 Activities of the Newly Synthesized 2-Substituted Thienopyridines. Curr. Org. Chem. 2009, 13, 1654–1663. [Google Scholar] [CrossRef]
  24. Binsaleh, N.K.; Wigley, C.A.; Whitehead, K.A.; van Rensburg, M.; Reynisson, J.; Pilkington, L.I.; Barker, D.; Jones, S.; Dempsey-Hibbert, N.C. Thieno[2,3-b]pyridine derivatives are potent anti-platelet drugs, inhibiting platelet activation, aggregation and showing synergy with aspirin. Eur. J. Med. Chem. 2018, 143, 1997–2004. [Google Scholar] [CrossRef] [Green Version]
  25. Amorim, R.; de Meneses, M.D.F.; Borges, J.C.; da Silva Pinheiro, L.C.; Caldas, L.A.; Cirne-Santos, C.C.; de Mello, M.V.P.; de Souza, A.M.T.; Castro, H.C.; de Palmer Paixão, I.C.N.; et al. Thieno[2,3-b]pyridine derivatives: A new class of antiviral drugs against Mayaro virus. Arch. Virol. 2017, 162, 1577–1587. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, N.Y.; Zuo, W.Q.; Xu, Y.; Gao, C.; Zeng, X.X.; Zhang, L.D.; You, X.Y.; Peng, C.T.; Shen, Y.; Yang, S.Y.; et al. Discovery and structure−activity relationships study of novel thieno[2,3-b]pyridine analogues as hepatitis C virus inhibitors. Bioorg. Med. Chem. Lett. 2014, 24, 1581–1588. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, H.; Li, Y.; Wang, X.Y.; Wang, B.; He, H.Y.; Liu, J.Y.; Xiang, M.L.; He, J.; Wu, X.H.; Yang, L. Synthesis, preliminary structure-activity relationships, and in vitro biological evaluation of 6-aryl-3-amino-thieno[2,3-b]pyridine derivatives as potential anti-inflammatory agents. Bioorg. Med. Chem. Lett. 2013, 23, 2349–2352. [Google Scholar] [CrossRef] [PubMed]
  28. Ajmal, R.B.S. Biological Activity of Pyrimidine Derivativies: A Review. Org. Med. Chem. IJ. 2017, 2, 555581. [Google Scholar] [CrossRef]
  29. Patil, S.B. Biological and medicinal significance of pyrimidines: A review. Int. J. Pharm Sci. Res. 2018, 9, 44–52. [Google Scholar] [CrossRef]
  30. Kaur, H.; Machado, M.; de Kock, C.; Smith, P.; Chibale, K.; Prudêncio, M.; Singh, K. Primaquine-pyrimidine hybrids: Synthesis and dual-stage antiplasmodial activity. Eur. J. Med. Chem. 2015, 101, 266–273. [Google Scholar] [CrossRef]
  31. Barakat, A.; Soliman, S.M.; Al-Majid, A.M.; Lotfy, G.; Ghabbour, H.A.; Fun, H.K.; Yousuf, S.; Choudhary, M.I.; Wadood, A. Synthesis and structure investigation of novel pyrimidine-2,4,6-trione derivatives of highly potential biological activity as anti-diabetic agent. J. Mol. Struct. 2015, 1098, 365–376. [Google Scholar] [CrossRef]
  32. Su, L.; Li, J.; Zhou, Z.; Huang, D.; Zhang, Y.; Pei, H.; Guo, W.; Wu, H.; Wang, X.; Liu, M.; et al. Corrigendum to Design, synthesis and evaluation of hybrid of tetrahydrocarbazole with 2,4-diaminopyrimidine scaffold as antibacterial agents. Eur. J. Med. Chem. 2019, 168, 385. [Google Scholar] [CrossRef] [PubMed]
  33. Bassyouni, F.; Tarek, M.; Salama, A.; Ibrahim, B.; Salah El Dine, S.; Yassin, N.; Hassanein, A.; Moharam, M.; Abdel-Rehim, M. Promising Antidiabetic and Antimicrobial Agents Based on Fused Pyrimidine Derivatives: Molecular Modeling and Biological Evaluation with Histopathological Effect. Molecules 2021, 26, 2370. [Google Scholar] [CrossRef] [PubMed]
  34. Ayati, A.; Moghimi, S.; Toolabi, M.; Foroumadi, A. Pyrimidine-based EGFR TK Inhibitors in Targeted Cancer Therapy. Eur. J. Med. Chem. 2021, 221, 113523. [Google Scholar] [CrossRef]
  35. Tylińska, B.; Wiatrak, B.; Czyżnikowska, Ż.; Cieśla-Niechwiadowicz, A.; Gębarowska, E.; Janicka-Kłos, A. Novel Pyrimidine Derivatives as Potential Anticancer Agents: Synthesis, Biological Evaluation and Molecular Docking Study. Int. J. Mol. Sci. 2021, 22, 3825. [Google Scholar] [CrossRef] [PubMed]
  36. Mahapatra, A.; Prasad, T.; Sharma, T. Pyrimidine: A review on anticancer activity with key emphasis on SAR. Futur. J. Pharm. Sci. 2021, 7, 123. [Google Scholar] [CrossRef]
  37. Kim, Y.; Kim, M.; Park, M.; Tae, J.; Baek, D.J.; Park, K.D.; Choo, H. Synthesis of novel dihydropyridothienopyrimidin-4, 9-dione derivatives. Molecules 2015, 20, 5074–5084. [Google Scholar] [CrossRef] [Green Version]
  38. Mohi El-Deen, E.M.; Abd El-Meguid, E.A.; Karam, E.A.; Nossier, E.S.; Ahmed, M.F. Synthesis and Biological Evaluation of New Pyridothienopyrimidine Derivatives as Antibacterial Agents and Escherichia coli Topoisomerase II Inhibitors. Antibiotics 2020, 9, 695. [Google Scholar] [CrossRef]
  39. Sanad, S.M.H.; Mekky, A.E.M. New pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4(3H)-one hybrids linked to arene units: Synthesis of potential MRSA, VRE, and COX-2 inhibitors. Can. J. Chem. 2021, 99, 900–909. [Google Scholar] [CrossRef]
  40. Sirakanyan, S.N.; Spinelli, D.; Geronikaki, A.; Hakobyan, E.K.; Sahakyan, H.; Arabyan, E.; Zakaryan, H.; Nersesyan, L.E.; Aharonyan, A.S.; Danielyan, I.S.; et al. Synthesis, Antitumor Activity, and Docking Analysis of New Pyrido[3′,2′:4,5]furo(thieno)[3,2-d]pyrimidin-8-amines. Molecules 2019, 24, 3952. [Google Scholar] [CrossRef] [Green Version]
  41. Kang, M.A.; Kim, M.; Kim, J.Y.; Shin, Y.; Song, J.; Jeong, J. A novel pyrido-thieno-pyrimidine derivative activates p53 through induction of phosphorylation and acetylation in colorectal cancer cells. Int. J. Oncol. 2015, 46, 342–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Aziz, Y.M.A.; Said, M.M.; El Shihawy, H.A.; Abouzid, K.A. Discovery of novel tricyclic pyrido [3′, 2′: 4, 5] thieno [3, 2-d] pyrimidin-4-amine derivatives as VEGFR-2 inhibitors. Bioorg. Chem. 2015, 60, 1–12. [Google Scholar] [CrossRef] [PubMed]
  43. Loidreau, Y.; Deau, E.; Marchand, P.; Nourrisson, M.-R.; Logé, C.; Coadou, G.; Loaëc, N.; Meijer, L.; Besson, T. Synthesis and molecular modelling studies of 8-arylpyrido[3′,2′:4,5]thieno[3,2-d]pyrimidin-4-amines as multitarget Ser/Thr kinases inhibitors. Eur. J. Med. Chem. 2015, 92, 124. [Google Scholar] [CrossRef]
  44. Al-Ghorbani, M.; Bushra Begum, A.; Zabiulla, S.; Mamatha, S.V.; Ara Khanum, S. Piperazine and morpholine: Synthetic preview and pharmaceutical applications. J. Chem. Pharm. Res. 2015, 7, 281–301. [Google Scholar] [CrossRef]
  45. Sepsey Für, C.; Riszter, G.; SzigetvárI, Á.; Dékány, M.; Keglevich, G.; HazaI, L.; BölcskeI, H. Novel Ring Systems: Spiro[Cycloalkane] Derivatives of Triazolo- and Tetrazolo-Pyridazines. Molecules 2021, 26, 2140. [Google Scholar] [CrossRef]
  46. Zheng, Y.; Tice, C.M.; Suresh, B.; Singh, S.B. The use of spirocyclic scaffolds in drug discovery. Bioorganic Med. Chem. Lett. 2014, 24, 3673–3682. [Google Scholar] [CrossRef] [Green Version]
  47. Delost, M.D.; Smith, D.T.; Anderson, B.J.; Njardarson, J.T. From Oxiranes to Oligomers: Architectures of U.S. FDA Approved Pharmaceuticals Containing Oxygen Heterocycles. J. Med. Chem. 2018, 61, 10996–11020. [Google Scholar] [CrossRef]
  48. Chauhan, R.; Saini, P.; Choudhary, R.; Rani, S. A Review on Pharmacological Profile of Ethanamide and their Derivatives. Int. J. Pharm. Sci. Rev. Res. 2020, 64, 162–170. [Google Scholar] [CrossRef]
  49. Olayioye, M.A.; Neve, R.M.; Lane, H.A.; Hynes, N.E. The ErbB signaling network: Receptor heterodimerization in development and cancer. EMBO J. 2000, 19, 3159–3167. [Google Scholar] [CrossRef] [Green Version]
  50. de Castro Barbosa, M.L.; Lima, L.M.; Tesch, R.; Sant’ Anna, C.M.R.; Totzke, F.; Kubbutat, M.H.; Schächtele, C.; Laufer, S.; Barreiro, E.J. Novel 2-chloro-4-anilino-quinazoline derivatives as EGFR and VEGFR-2 dual inhibitors. Eur. J. Med. Chem. 2014, 71, 1–14. [Google Scholar] [CrossRef]
  51. Elshaier, Y.A.; Shaaban, M.A.; Abd El Hamid, M.K.; Abdelrahman, M.H.; Abou-Salim, M.A.; Elgazwi, S.M.; Halaweish, F. Design and synthesis of pyrazolo [3, 4-d] pyrimidines: Nitric oxide releasing compounds targeting hepatocellular carcinoma. Bioorg. Med. Chem. 2017, 25, 2956–2970. [Google Scholar] [CrossRef] [PubMed]
  52. Chang, J.; Ren, H.; Zhao, M.; Chong, Y.; Zhao, W.; He, Y.; Zhao, Y.; Zhang, H.; Qi, C. Development of a series of novel 4-anlinoquinazoline derivatives possessing quinazoline skeleton: Design, synthesis, EGFR kinase inhibitory efficacy, and evaluation of anticancer activities in vitro. Eur. J. Med. Chem. 2017, 138, 669–688. [Google Scholar] [CrossRef] [PubMed]
  53. van Meerloo, J.; Kaspers, G.J.; Cloos, J. Cell sensitivity assays: The MTT assay. Methods Mol. Biol. 2011, 731, 237–245. [Google Scholar] [CrossRef] [PubMed]
  54. Aiebchun, T.; Mahalapbutr, P.; Auepattanapong, A.; Khaikate, O.; Seetaha, S.; Tabtimmai, L.; Kuhakarn, C.; Choowongkomon, K.; Rungrotmongkol, T. Identification of vinyl sulfone derivatives as egfr tyrosine kinase inhibitor: In vitro and in silico studies. Molecules 2021, 26, 2211. [Google Scholar] [CrossRef] [PubMed]
  55. Lyseng-Williamson, K.A. Erlotinib: A pharmacoeconomic review of its use in advanced non-small cell lung cancer. Pharmacoeconomics 2010, 28, 75–92. [Google Scholar] [CrossRef] [PubMed]
  56. Youssefyeh, R.D.; Brown, R.E.; Wilson, J.; Shah, U.; Jones, H.; Loev, B.; Khandwala, A.; Leibowitz, M.J.; Sonnino-Goldman, P. Pyrido [3′, 2′: 4, 5] thieno [3, 2-d]-N-triazines: A new series of orally active antiallergic agents. J. Med. Chem. 1984, 27, 1639–1643. [Google Scholar] [CrossRef]
  57. Wiegand, I.; Hilpert, K.; Hancock, R.E. Agar and broth dilution methods to determine the minimal inhibitory concentration (MIC) of antimicrobial substances. Nat. Protoc. 2008, 3, 163–175. [Google Scholar] [CrossRef]
  58. Mizutani, M.Y.; Takamatsu, Y.; Ichinose, T.; Nakamura, K.; Itai, A. Effective handling of induced-fit motion in flexible docking. Proteins Struct. Funct. Bioinform. 2006, 63, 878–891. [Google Scholar] [CrossRef]
  59. Stamos, J.; Sliwkowski, M.X.; Eigenbrot, C. Structure of the epidermal growth factor receptor kinase domain alone and in complex with a 4-anilinoquinazoline inhibitor. J. Biol. Chem. 2002, 277, 46265–46272. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Reported thienopyridines (IIII) and pyrimidines (IVVI) effective as antimicrobial and/or anticancer agents, and design of the new pyridothienopyrimidine compounds.
Figure 1. Reported thienopyridines (IIII) and pyrimidines (IVVI) effective as antimicrobial and/or anticancer agents, and design of the new pyridothienopyrimidine compounds.
Molecules 27 00803 g001
Scheme 1. Synthesis of pyridothienopyridine derivatives 2a,b5a,b.
Scheme 1. Synthesis of pyridothienopyridine derivatives 2a,b5a,b.
Molecules 27 00803 sch001
Scheme 2. Synthesis of 1′H-spirocycloalkane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine derivatives 6a–d–9a,b.
Scheme 2. Synthesis of 1′H-spirocycloalkane-1,2′-pyrido[3′,2′:4,5]thieno[3,2-d]pyrimidine derivatives 6a–d–9a,b.
Molecules 27 00803 sch002
Figure 2. Antibacterial activities (MIC in µg/mL) of the new pyridothienpyrimidine compounds 2a,b–9a,b and the reference antibiotic (amoxicillin trihydrate).
Figure 2. Antibacterial activities (MIC in µg/mL) of the new pyridothienpyrimidine compounds 2a,b–9a,b and the reference antibiotic (amoxicillin trihydrate).
Molecules 27 00803 g002
Figure 3. The antifungal activities (MIC in µg/mL) of the new pyridothienpyrimidine compounds 2a,b–9a,b and the reference drug (clotrimazole).
Figure 3. The antifungal activities (MIC in µg/mL) of the new pyridothienpyrimidine compounds 2a,b–9a,b and the reference drug (clotrimazole).
Molecules 27 00803 g003
Figure 4. The cytotoxic activity of the most potent compounds.
Figure 4. The cytotoxic activity of the most potent compounds.
Molecules 27 00803 g004
Figure 5. 3D representation of the superimposition of the co-crystallized ligand (purple) and the docking pose (dark grey) of ERL at the active site of the EGFR enzyme.
Figure 5. 3D representation of the superimposition of the co-crystallized ligand (purple) and the docking pose (dark grey) of ERL at the active site of the EGFR enzyme.
Molecules 27 00803 g005
Figure 6. (a) 2D interactions and (b) 3D interactions of erlotinib within the EGFR kinase domain’s active site.
Figure 6. (a) 2D interactions and (b) 3D interactions of erlotinib within the EGFR kinase domain’s active site.
Molecules 27 00803 g006
Figure 7. (a) 2D interactions and (b) 3D interactions of compound 9b within the EGFR kinase domain’s active site.
Figure 7. (a) 2D interactions and (b) 3D interactions of compound 9b within the EGFR kinase domain’s active site.
Molecules 27 00803 g007
Figure 8. (a) 2D interactions and (b) 3D interactions of compound 5a within the EGFR kinase domain’s active site.
Figure 8. (a) 2D interactions and (b) 3D interactions of compound 5a within the EGFR kinase domain’s active site.
Molecules 27 00803 g008
Figure 9. (a) 2D interactions and (b) 3D interactions of compound 3a within the EGFR kinase domain’s active site.
Figure 9. (a) 2D interactions and (b) 3D interactions of compound 3a within the EGFR kinase domain’s active site.
Molecules 27 00803 g009
Table 1. In vitro antibacterial activities of the synthesized compounds expressed as MICs (μg/mL) against the tested pathogenic bacteria.
Table 1. In vitro antibacterial activities of the synthesized compounds expressed as MICs (μg/mL) against the tested pathogenic bacteria.
Compd. No.Gram +ve BacteriaGram −ve Bacteria
S. aureusB. subtilisB. cereusE. coliS. typhimurium
2aNANA64NANA
2b641286412864
3a448816
3b328321632
4a484816
4b812816128128
5a8816816
5b88163232
6a321683216
6b4816816
6c448816
6d83283232
7a4NA8816
7b64646464128
7c323283232
7d464643232
8a81683232
8b4816816
9a163283216
9b448416
Amoxicillin4816816
NA = No Activity (MIC ˃ 128 µg/mL).
Table 2. In vitro antifungal activities of the synthesized compounds expressed as MICs (μg/mL) against the tested pathogenic yeasts and fungi.
Table 2. In vitro antifungal activities of the synthesized compounds expressed as MICs (μg/mL) against the tested pathogenic yeasts and fungi.
Compd. No.YeastsFungi
C. albicansC. tropicalsS. cerevisiaeA. flavusA. niger
2a12812816NANA
2b6432323232
3a168888
3b3216643216
4a1616888
4b81681616
5a1688816
5b6416163264
6a6416326432
6b888168
6c1641648
6d6464323216
7a161681616
7b64643264128
7c1688416
7d3264326464
8a6416161616
8b8161688
9a1632321664
9b44888
Clotrimazole168888
NA = No Activity (MIC ˃ 128 µg/mL).
Table 3. Cytotoxic activities (IC50 µM) of the new compounds and doxorubicin against HepG2, MCF7 and WISH cells.
Table 3. Cytotoxic activities (IC50 µM) of the new compounds and doxorubicin against HepG2, MCF7 and WISH cells.
Compd. No.HepG2 (µM)MCF7 (µM)WISH
2a56.57 ± 3.6464.34 ± 2.91
2b33.21 ± 1.7942.39 ± 2.24
3a2.31 ± 0.357.24 ± 0.64416.83 ± 15.17
3b11.34 ± 0.6524.72 ± 1.32
4a2.99 ± 0.1515.42 ± 0.45460.23 ± 11.08
4b36.52 ± 1.8243.27 ± 2.28
5a1.99 ± 0.092.79 ± 0.18408.48 ± 15.93
5b10.16 ± 0.2921.06 ± 1.16
6a52.18 ± 1.4577.41 ± 3.62
6b2.75 ± 0.139.89 ± 0.55394.98 ± 10.20
6c12.11 ± 0.3322.24 ± 0.67
6d4.45 ± 0.2221.67 ± 0.70
7a26.05 ± 1.8928.11 ± 0.92
7b39.74 ± 1.8941.62 ± 2.20
7c10.35 ± 0.3120.9 ± 0.93
7d23.25 ± 0.3526.55 ± 1.62
8a6.78 ± 0.7320.88 ± 1.46
8b2.79 ± 0.0813.54 ± 0.76401.37± 17.32
9a4.88 ± 0.6523.56 ± 1.24
9b1.17 ± 0.091.52 ± 0.08417.55 ± 14.1
Doxorubicin2.85 ± 0.213.58 ± 0.33432.10 ± 19.30
Table 4. In vitro enzymatic inhibitory activity against EGFR kinase.
Table 4. In vitro enzymatic inhibitory activity against EGFR kinase.
Compound No.EGFR
IC50 (nM)
3a17.29 ± 0.24
4a53.57 ± 0.41
5a9.66 ± 0.08
6b53.19 ± 0.46
8b38.44 ± 0.25
9b7.27 ± 0.11
Erlotinib27.01 ± 0.16
Table 5. Molecular docking results of the most active pyridothienopyrimidine compounds.
Table 5. Molecular docking results of the most active pyridothienopyrimidine compounds.
CompoundS (kcal/mol)Amino AcidsInteracting GroupsType of BondLength (Å)
3a–11.42Val702N (Pyrimidine)H-bond acceptor4.12
Lys721N (Pyridine)H-bond acceptor3.25
Met769Cl (Pyrimidine)Halogen bond3.47
Leu768Cl (Pyrimidine)Halogen bond4.16
Thr830Cl (Pyridine)Halogen bond3.26
Met742Cl (Pyridine)Halogen bond4.13
Thr766Cl (Pyridine)Halogen bond3.03
Thr830Sσ-hole bond3.79
4a–8.94Lys721ClHalogen bond3.69
Cys751O (C=O)σ-hole bond3.78
Thr766Sσ-hole bond4.21
Met769S and N (Pyridine)H-bond acceptor3.41/3.55
Leu768N (Pyridine)H-bond acceptor3.84
5a–11.48Asp831NH+Ionic interaction3.71
Lys721O (Morpholine)H-bond acceptor3.51
Cys773O (C=O)σ-hole bond3.46
Asp776Sσ-hole bond4.18
6b–10.06Leu820NHH-bond acceptor3.78
Cys751O (C=O)H-bond acceptor3.49
Gln767Sσ-hole bond3.12
Thr766Sσ-hole bond3.87
Met769SH-bond acceptor3.83
8b–9.11Leu694S (Thiophene)H-bond acceptor4.38
Leu694S (Side chain)σ-hole bond3.79
Thr766N (Pyridine)H-bond acceptor3.71
9b–12.01Met769S (thiophene)H-bond acceptor4.09
Leu768S (thiophene)H-bond acceptor4.37
Leu820S (thiophene)H-bond acceptor4.47
Thr766S (Side chain)σ-hole bond4.20
Leu820N and NH (Pyrimidine)H-bond acceptor3.73/3.74
Val702O (Oxirane)H-bond acceptor3.64
erlotinib–10.48Leu768 N (Pyrimidine)H-bond acceptor3.64
Met769N (Pyrimidine)H-bond acceptor2.70
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/molecules27030803

AMA Style

Mohi El-Deen EM, Anwar MM, El-Gwaad AAA, Karam EA, El-Ashrey MK, Kassab RR. Novel Pyridothienopyrimidine Derivatives: Design, Synthesis and Biological Evaluation as Antimicrobial and Anticancer Agents. Molecules. 2022; 27(3):803. https://doi.org/10.3390/molecules27030803

Chicago/Turabian Style

Mohi El-Deen, Eman M., Manal M. Anwar, Amina A. Abd El-Gwaad, Eman A. Karam, Mohamed K. El-Ashrey, and Rafika R. Kassab. 2022. "Novel Pyridothienopyrimidine Derivatives: Design, Synthesis and Biological Evaluation as Antimicrobial and Anticancer Agents" Molecules 27, no. 3: 803. https://doi.org/10.3390/molecules27030803

Article Metrics

Back to TopTop