Next Article in Journal
Amphiphilic Cell-Penetrating Peptides Containing Arginine and Hydrophobic Residues as Protein Delivery Agents
Next Article in Special Issue
In Vitro and In Vivo Effects of Ulvan Polysaccharides from Ulva rigida
Previous Article in Journal
Drug-Utilization, Healthcare Facilities Accesses and Costs of the First Generation of JAK Inhibitors in Rheumatoid Arthritis
Previous Article in Special Issue
Indoline-5-Sulfonamides: A Role of the Core in Inhibition of Cancer-Related Carbonic Anhydrases, Antiproliferative Activity and Circumventing of Multidrug Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Pot Synthesis of 1-Thia-4-azaspiro[4.4/5]alkan-3-ones via Schiff Base: Design, Synthesis, and Apoptotic Antiproliferative Properties of Dual EGFR/BRAFV600E Inhibitors

by
Lamya H. Al-Wahaibi
1,
Essmat M. El-Sheref
2,
Mohamed M. Hammouda
3,4 and
Bahaa G. M. Youssif
5,*
1
Department of Chemistry, College of Sciences, Princess Nourah Bint Abdulrahman University, Riyadh 11564, Saudi Arabia
2
Chemistry Department, Faculty of Science, Minia University, El Minia 61519, Egypt
3
Department of Chemistry, College of Science and Humanities in Al-Kharj, Prince Sattam Bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
4
Chemistry Department, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
5
Pharmaceutical Organic Chemistry Department, Faculty of Pharmacy, Assiut University, Assiut 71526, Egypt
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2023, 16(3), 467; https://doi.org/10.3390/ph16030467
Submission received: 28 February 2023 / Revised: 13 March 2023 / Accepted: 20 March 2023 / Published: 22 March 2023
(This article belongs to the Special Issue Novel Anti-proliferative Agents)

Abstract

:
In this investigation, novel 4-((quinolin-4-yl)amino)-thia-azaspiro[4.4/5]alkan-3-ones were synthesized via interactions between 4-(2-cyclodenehydrazinyl)quinolin-2(1H)-one and thioglycolic acid catalyzed by thioglycolic acid. We prepared a new family of spiro-thiazolidinone derivatives in a one-step reaction with excellent yields (67–79%). The various NMR, mass spectra, and elemental analyses verified the structures of all the newly obtained compounds. The antiproliferative effects of 6ae, 7a, and 7b against four cancer cells were investigated. The most effective antiproliferative compounds were 6b, 6e, and 7b. Compounds 6b and 7b inhibited EGFR with IC50 values of 84 and 78 nM, respectively. Additionally, 6b and 7b were the most effective inhibitors of BRAFV600E (IC50 = 108 and 96 nM, respectively) and cancer cell proliferation (GI50 = 35 and 32 nM against four cancer cell lines, respectively). Finally, the apoptosis assay results revealed that compounds 6b and 7b had dual EGFR/BRAFV600E inhibitory properties and showed promising antiproliferative and apoptotic activity.

1. Introduction

The development of new drugs in anti-cancer research depends on a better understanding of druggable targets. According to this strategy, changing particular cancer biomarkers will produce beneficial therapeutic effects [1]. Selective anti-cancer medications must be more effective at destroying tumors while having fewer side effects on normal cells [2]. Single-target therapy has been recognized as causing chemotherapeutic resistance [3]. Recently, combination therapy (drug cocktails combining two or more therapeutic agents with different modes of action and non-overlapping toxicities) was authorized as an alternative to single-target chemotherapy for cancer [4]. Despite the possibility of additive and synergistic effects, combination therapy frequently results in unexpected adverse effects, such as increased toxicity. Alternatives to combination therapy include drugs with two or more targets, a reduced risk of drug interactions, enhanced pharmacokinetics (PK), and improved safety profiles [5]. Additionally, a dual-target kinase may inhibit drug interactions, harmful off-target impacts, poor patient compliance, and elevated costs of production [5].
Combining tyrosine kinase (TK) and BRAF inhibitors has proven useful in preventing tumor growth and minimizes resistance in clinical trials. Combining vemurafenib and EGFR inhibitors in thyroid carcinoma may help overcome resistance to BRAF inhibitors [6]. This combination has also been effective in BRAFV600E colorectal cancer [7]. Additionally, several in vitro substances have been developed, including EGFR/VEGFR-2 and BRAF, which contain the essential pharmacophoric groups necessary to suppress tyrosine kinase [8,9]. In summary, dual/multi-targeted kinase inhibitors can be turned into effective anti-cancer medications [10].
In medicinal chemistry, nitrogen heterocycles are crucial structural components. The chemical and biological uses of quinoline, one of many heterocyclic compounds, have been studied by numerous research groups [11,12]. The FDA recently approved quinazoline derivatives, such as gefitinib and erlotinib (Figure 1), as EGFR inhibitors for treating non-small cell breast and lung cancers [13,14,15]. In addition, Thr766 in the EGFR pocket forms a water-mediated hydrogen bond with the nitrogen atom at position 3 of the quinazoline core [16,17]. New quinoline derivatives, such as pelitinib and neratinib, which are potent EGFR inhibitors, were developed through the bioisosteric replacement of the quinazoline core with a quinoline ring (Figure 1) [18,19,20,21,22]. The bioisosteric substitution did not require water molecules to facilitate the interaction with amino acid residue Thr766 [23].
We previously reported synthesizing two novel series of quinoline-2-one-based derivatives as potential apoptotic antiproliferative agents targeting the EGFR inhibitory pathway [24]. Compounds I and II (Figure 1) inhibited EGFR effectively, with IC50 values of 0.18 and 0.09 µM, respectively. Furthermore, the two compounds significantly increased the apoptotic markers caspase-3, caspase-8, and the Bax levels while decreasing antiapoptotic Bcl2. In another study [25], compound III, a quinoline-based derivative, was developed as a dual EGFR and BRAFV600E inhibitor with IC50 values of 1.30 and 3.8 µM, respectively.
The Spiro scaffold is another building material with potential applications in medicinal chemistry. Chemists have introduced a ring for rigidity when designing new drugs to potentially reduce the entropic penalties upon binding to target proteins. Conformational restriction can also be achieved by spiro ring fusion. They are frequently used in the design and discovery of drugs due to their inherent three-dimensionality and novel structural characteristics. Previously, spiro ring systems were efficiently integrated into enzyme inhibitors, with kinases topping the list, as well as protein–protein interaction inhibitors [26,27,28].
On the other hand, the thiazole moiety has piqued researchers’ interest due to its robust biological properties [29,30]. Regarding anti-cancer activity, thiazole and its derivatives rank among the most active compounds [31,32,33,34,35,36]. Furthermore, several clinically available anti-cancer drugs contain thiazole-containing compounds, such as dabrafenib (IV, Figure 2) (BRAF inhibitor) [37].
Abdel-Maksoud et al. also developed a novel series of thiazole-based derivatives for testing as BRAFV600E inhibitors. Compound V (Figure 2) exhibited the most potent antiproliferative activity, with potent BRAFV600E inhibitory activity and an IC50 value of 0.05 µM [38]. We recently reported the synthesis of novel thiazole-based derivatives that act as dual EGFR and BRAFV600E inhibitors [39]. Compound VI (Figure 2) demonstrated significant antiproliferative activity against EGFR and BRAFV600E, with IC50 values of 74 and 107 nM, respectively.
Following our previous study on dual targeting strategies and motivated by the findings above [40,41,42], we synthesized new spiro-compounds that combine quinolinone, thiazolidinone, and spiro-cyclic in a single molecule via the reaction of thioglycolic acid and 4-(2-cyclodenehydrazinyl)4uinoline-2(1H)-one. As a result, two new compounds, 6ae (Scaffold A) and 7a and 7b (Scaffold B), were developed to generate new potent antiproliferative agents targeting EGFR and/or BRAFV600E, as shown in Figure 3.
The new compounds underwent a cell viability assay test to determine their influence on the vitality of normal cell lines. They were also tested against a panel of four cancer cell lines as antiproliferative agents. Furthermore, the most active compounds were tested for antiproliferative activity against EGFR and BRAFV600E. Finally, we tested the most active compounds for apoptotic potential against caspase 3, caspase 8, Bax, and the antiapoptotic Bcl2.

2. Results and Discussion

2.1. Chemistry

We aimed to develop a novel series of spiro-compounds, namely 6ae, 7a and 7b. Scheme 1 depicts the steps taken to obtain our target 4-((6-substituted-2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4-azaspiroalkan-3-ones (6ae) in high yields via the interaction between 4-(2-cyclodenehydrazinyl)5uinoline-2(1H)-ones (5ae) [43,44,45] and thioglycolic acid refluxed under dry benzene with a free catalyst at a molar ratio of 1:20 for 20 h.
Reagents and reaction conditions: (a) POCl3/stirring for 1 h, 70 °C; (b) AcOH/H2O/refluxing overnight; (c) Hydrazine/EtOH/refluxing for 3 h; (d) Cyclic ketone/EtOH/refluxing for 3 h; I Thioglycolic acid/Benzene/refluxing for 24 h.
The structures of our obtained products, 6a–e, were confirmed using NMR, mass spectrometry, and elemental analysis. All of the spectral data show that the acquired molecular formula for 6a–e comprises one molecule from compounds 5ae and one molecule of thioglycolic acid, with the elimination of the H2O molecule. Compound 6b [4-((2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4-azaspiro-[4.5]decan-3-one], with the chemical formula C17H19N3O2S and a detectable molecular ion peak at m/z = 329, was used as a representative example. Moreover, the 1H NMR spectra of compound 6b revealed two broad singlet signals at δH = 11.01 and 10.07 ppm, respectively, indicative of NH-1 and exo-NH-4b. Two singlet signals at δH = 6.11 and 3.74 ppm were assigned as H-3 and H-2′, respectively, in addition to quinolinone-CH between 7.97 and 7.12 ppm (m, 4H). The 13C NMR spectra for compound 6b had four downfield-lying lines at δC = 167.56 (C-3′), 162.85 (C-2), 151.76 (C-4), and 139.06 (C-8a) ppm, respectively. C-5 resonated at δC = 61.27 ppm and was characteristic of a spiro-structure.
The possibility of an isomeric structure in compound 6b′ (Figure 4) was ruled out by 1H NMR, which revealed an NH chemical shift at 10.07 ppm due to hydrogen bonding with the carbonyl group (C-3′). Furthermore, the 15N NMR spectrum revealed a signal at δN = 131.7 ppm, which produced an HSQC correlation with the proton at δH = 10.07 ppm. An HMBC correlation with two protons at δH = 10.07 and 6.11 ppm, assigned as NH-4b and H-3, was also indicated (Figure 5). The latter applies to structure 6b but not 6b′. Furthermore, the two signals at δN = 142.1 ppm, designated as NH-1, show an HSQC correlation with the proton at H 6.11 ppm, designated as H-3. These correlations are not possible for structure 6b′.
We also investigated the reaction of thioglycolic acid with 4-(2-(2,2,6,6-tetramethylpiperidin-4-ylidene)hydrazinyl)quinolin-2(1H)-one (5f) and 4-(2-(1-benzylpiperidin-4-ylidene) hydrazine-yl)quinolin-2(1H)-one (5g). The reaction followed the same pattern and produced similar products, 7a and 7b, in high yields (Scheme 2).
For example, consider compound 7b, which was designated as 8-benzyl-4-((2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4,8-diazaspiro[4.5]decan-3-one with the molecular formula C23H24N4O2S (Figure 6, Table 1). The 1H NMR of compound 7b revealed that H-3 was distinctive as the vinylic singlet at δH 6.08 ppm, with its attached carbon appearing at δC 92.33 ppm. Furthermore, H-3 exhibits an HMBC correlation with both protonated nitrogens at δN 141.9 and 132.0 ppm. The latter could be N-1 and N-4b, although it is unclear which was which. Furthermore, N-1 exhibits an HMBC correlation with the proton resonance at δH 7.28 ppm, implying that H-8 is present. Of the two carbons that exhibit an HSQC correlation with this proton, the one at δC 115.45 ppm exhibits an HMBC correlation with the protons in the quinoline ring, designated as C-8. The only visible 1H doublet in the aromatic region appears at δH 7.98 ppm and is assigned to H-5, whose attached carbon appears at δC 122.44 ppm. H-5 establishes a COSY correlation with the dd at δH 7.12 ppm, designated as H-6. The attached carbon appears at δC 120.31 ppm in H-6, establishing a COSY correlation with the other dd, at δH 7.46, designated as H-7. Attached carbon appears at δC 130.13 ppm in H-7. In turn, H-7 establishes a COSY correlation with H-8. The carbon at δC 139.23 establishes an HMBC correlation with H-5 and H-7, designated as C-8a. The carbon at δC 112.26 establishes a strong HMBC correlation with H-3, NH-4b, H-6, and H-8, designated as C-4a. These six HMBC correlations are three-bond correlations. The carbon at δC 162.90 establishes an HMBC correlation with H-3 and (weakly) NH-4b, designated as C-2. The carbon at δC 149.51 establishes an HMBC correlation with H-3, NH-4b, H-5, and (weakly) H-8, designated as C-4. This arrangement constituted the quinolinone structure.
Furthermore, H-8″ is distinctive as the 2H singlet at δH 3.56, whose attached carbon appears at δC 61.36 ppm., and establishes an HMBC correlation with the signal at δC 128.73 ppm, designated as C-o. Both C-o and C-m establish HSQC correlations with the m-signal at δH 7.35 ppm. Furthermore, the carbon at δC 126.93 establishes a COSY correlation with both H-o and H-m, and HSQC correlations with the signal at 7.28 ppm. This carbon must be C-p, as shown in Table 1.
Compounds 6ae were synthesized by cyclizing the corresponding hydrazones 5ae with thioglycolic acid according to the suggested mechanism in Scheme 3. Compounds 5ac were catalyzed with thioglycolic acid to form tertiary carbocation 8 and a nucleophilic addition of HS-thioglycolic, followed by rearrangement with the loss of an H+-proton to produce the intermediate 9. Nucleophilic attacks from NH-2 to the carbonyl group eliminated an H2O molecule, yielding the corresponding products 6a–e via intermediates 9 and 10. Compounds 7a and 7b were also formed via a similar mechanism.

2.2. Biology

2.2.1. Cell Viability Assay

The epithelial (MCF-10A) cell line of a normal human mammary gland was used to test the viability of the new substances. Compounds 6ae, 7a, and 7b were incubated on MCF-10A cells for four days before being tested for viability using the MTT assay [46,47]. According to Table 2, none of the tested substances revealed cytotoxic impacts, and the cell viability for the compounds tested at 50 µM was >89%.

2.2.2. Antiproliferative Assay

The antiproliferative role of 6ae, 7a, and 7b was tested against the four human cancer cell lines, namely MCF-7 (breast cancer cell line), Panc-1 (pancreatic cancer cell line), A-549 (lung cancer cell line), and HT-29 (colon cancer cell line), using the MTT assay and erlotinib as the reference drug [48,49]. Table 2 shows the median inhibitory concentration of each tested compound (IC50).
Overall, the recently screened derivatives 6ae, 7a, and 7b showed encouraging antiproliferative action against the four cancer cell lines tested, with an average IC50 (GI50) ranging between 32 and 81 nM compared to the reference erlotinib, which had a GI50 of 33 nM. Compounds 6b, 6e, and 7b had the highest antiproliferative activity, with GI50 values of 35, 40, and 32 nM, respectively. Compound 7b (R = Ph-CH2, R1 = R2 = H, Scaffold B) was the most effective derivative, with a GI50 value of 32 nM. It was comparable to, and even more effective than, erlotinib (GI50 = 33 nM) against the MCF-7 cell line, as shown in Table 3. The other Scaffold B compound 7a (R = H, R1 = R2= CH3) was the least effective derivative, with GI50 values of 81 nM. It was 2.5-fold less potent than 7b, suggesting the significance of the N-benzyl piperidine moiety for antiproliferative activity.
Compounds 6b (R = H, n = 1, Scaffold A) and 6e (R = OCH3, n = 1, Scaffold A) were the second and third most active compounds, with GI50 values of 35 and 40 nM, respectively. They were 1.1- and 1.25-fold less potent than 7b. Additionally, compound 6c (R = CH3, n = 1, Scaffold A) showed modest antiproliferative activity with a GI50 value of 73 nM against the four cancer cell lines tested. Compound 6c was 2.1- and 1.8-folds less potent than 6b and 6e, respectively.
Compound 6a (R = H, n = 0, Scaffold A) was 1.3-fold less potent than compound 6b (R = H, n = 1, Scaffold A) against the four cancer cell lines, with a GI50 value of 45 nM. The same pattern was observed when comparing compounds 6d (R = OCH3, n = 0, Scaffold A) and 6e (R = OCH3, n = 1, Scaffold A). Compound 6d had a GI50 of 52 nM, whereas 6e had a GI50 of 40 nM. These results showed that the spiro moiety’s ring size and the type of substitution at the quinoline moiety’s 6-position significantly impacted the antiproliferative function. Regarding the ring size, the cyclohexyl ring moiety was more tolerated for antiproliferative activity than the cyclopentyl ring moiety. Furthermore, the nature of substitution at the quinoline 6-position was associated with an increase in the antiproliferative activity, in the order H > OCH3 > CH3.

2.2.3. EGFR Inhibitory Assay

As potential targets for their antiproliferative activity, compounds 6ae, 7a, and 7b were tested for EGFR inhibitory activity [50]. Table 3 displays these results as IC50 values. The outcomes of this test align with the antiproliferative test, where the two most effective antiproliferative derivatives, compounds 7b (R = Ph-CH2, R1 = R2 = H, Scaffold B) and 6b (R = H, n = 1, Scaffold A), were the most effective EGFR inhibitors, with IC values of 78 ± 05 and 84 ± 06 nM, respectively. They were also equipotent to the reference erlotinib (IC50 = 80 ± 05).
Compounds 6a (R = H, n = 0, Scaffold A), 6d (R = OCH3, n = 0, Scaffold A), and 6e (R = OCH3, n = 1, Scaffold A) displayed significant anti-EGFR activity with IC50 values of 97 ± 07, 104 ± 08, and 92 ± 07 nM, respectively. They were 1.2- to 1.3-folds less potent than erlotinib.
Once again, compounds 6c (R = CH3, n = 1, Scaffold A) and 7a (R = H, R1 = R2 = CH3, Scaffold B) had the lowest activity as EGFR inhibitors, with IC50 values of 135 ± 11 and 149 ± 12 nM, respectively. These findings demonstrated that compounds 6b and 7b are workable antiproliferative candidates with substantial EGFR inhibitory properties.

2.2.4. BRAFV600E Inhibitory Assay

The anti-BRAFV600E activity of compounds 6ae, 7a, and 7b was also investigated in vitro [51], using erlotinib as a reference compound; the results are displayed in Table 3. According to the enzyme testing results, the investigated compounds had moderate to good BRAFV600E inhibitory activity, with IC50 values between 96 and 187 nM. In every instance, the compounds were less effective than the standard drug erlotinib (IC50 = 60 nM).
Table 3 shows that the most potent EGFR inhibitors, compounds 7b (R = Ph-CH2, R1 = R2 = H, Scaffold B) and 6b (R = H, n = 1, Scaffold A), were also the most potent BRAFV600E inhibitors, with IC values of 96 ± 8 and 108 ± 9 nM, respectively. Compounds 7b and 6b were roughly equivalent to erlotinib as antiproliferative agents; however, as BRAFV600E inhibitors, they were 1.6- and 1.8-fold less potent. Compounds 6a, 6c, 6d, 6e, and 7a showed weak to moderate anti-BRAF activity with IC50 values ranging between 129 and 187 nM, as shown in Table 3. The in vitro assay findings revealed that compounds 6b and 7b are potent antiproliferative agents that can act as dual EGFR/BRAFV600E inhibitors.

2.2.5. Apoptotic Markers Activation Assay

Numerous biochemical and morphological processes are involved in apoptosis, also called programmed cell death [52]. Antiapoptotic proteins such as Bcl-2 and Bc-W coexist with proapoptotic proteins such as Bad and Bax [53]. Proapoptotic proteins stimulate the release of cytochrome-c, whereas antiapoptotic proteins control apoptosis by inhibiting the release of cytochrome-c. The outer mitochondrial membrane becomes permeable when the ratio of proapoptotic proteins exceeds that of antiapoptotic proteins, setting off a series of events. The release of cytochrome c triggers both caspase-3 and caspase-9. Then, caspase-3 induces apoptosis by attacking several of the vital proteins that the cell requires [53].

Caspase 3 Activation Assay

Compounds 6b and 7b were tested as caspase-3 activators against the human pancreatic (Panc-1) cancer cell line [54]; the findings are shown in Table 4. Our results showed that derivatives 6b and 7b had a significant overexpression of caspase-3 protein levels (487.50 ± 4 and 544.50 ± 5 pg/mL, respectively) compared to the reference staurosporine (503.00 ± 4 pg/mL). The highest active substance, 7b, caused caspase-3 protein overexpression (544.50 ± 5 pg/mL) in the Panc-1 cancer cell line, which was 8.5-fold higher than the control with untreated cells, and even higher than staurosporine. Compound 6b (487.50 ± 4 pg/mL) showed comparable caspase-3 activation to the reference staurosporine. The above findings suggest that apoptosis may have contributed to the antiproliferative action of the tested compounds and caspase-3 overexpression.

Caspase-8, Bax and Bcl-2 Levels Assay

Compounds 6b and 7b were investigated for their impact on the Bacl-2, Bax, and caspase-8 levels against the Panc-1 cancer cell line, with staurosporine as a reference [55]. Our results are presented in Table 5.
Our findings demonstrated that the investigated substances significantly increased the Bax and caspase-8 levels compared to staurosporine. Compound 7b had the highest level of caspase-8 overexpression (1.90 ng/mL), followed by compound 6b (1.50 ng/mL) and the reference staurosporine (1.80 ng/mL). Additionally, 7b showed a comparable induction of Bax (295 pg/mL) to staurosporine (280 pg/mL), which was 37-fold higher than the control with untreated Panc-1 cancer cells. Finally, compound 7b caused the equipotent down-regulation of the Bcl-2 protein level (1.00 ng/mL), followed by compound 6b (1.30 ng/mL) in the Panc-1 cell line, and the reference staurosporine (1.10 ng/mL). The apoptosis assay results revealed that compounds 6b and 7b possess dual EGFR/BRAF inhibitory properties and show promising antiproliferative and apoptotic activity.

3. Conclusions

We designed and synthesized a small set of novel compounds, 4-((quinolin-4-yl)amino)thia-azaspiro[4.4/5]alkan-3-ones 6ae, 7a, and 7b, in order to develop new dual-targeting antiproliferative agents. The MTT assay was used to investigate the antiproliferative effects of the new compounds against a panel of four cancer cell lines. Compounds 6b, 6e, and 7b had higher antiproliferative activity, with GI50 values of 35, 40, and 32 nM, than erlotinib, which had a GI50 of 33 nM. The most potent compounds were then tested for EGFR and BRAFV600E inhibition. Our results showed that compounds 6b, 6e, and 7b were the most potent EGFR inhibitors, with IC50 values of 84, 92, and 78 nM, respectively. Furthermore, 6b, 6e, and 7b showed promising BRAFV600E inhibitory activity, with IC50 values of 108, 129, and 96 nM, respectively. Therefore, they can be considered effective antiproliferative agents that act as dual EGFR/BRAFV600E inhibitors. The most active derivative in Panc-1 cells, 7b, induces apoptosis via caspase 3 overexpression and an increased ratio of Bax/Bcl-2 genes compared to the control cells. In the future, more in vitro and in vivo studies, as well as chemical modifications, may be required to develop highly effective antiproliferative agents.

4. Experimental

4.1. Chemistry

General details: See Appendix SA (Supplementary File).
The 4-(2-substitutedhydrazinyl)-6-substituted-quinolin-2(1H)-one 5ag were prepared as reported [36,37,38], the thioglycolic acid (Aldrich) was used as received.
General procedure for preparation of compounds 6ae and 7a,b
The 5ag (1 mmol) and thioglycolic acid (20 mmol) were refluxed in 20 mL dray benzene in a round-bottom flask for 20 h. The reaction mixture was allowed to cool, and a white precipitate formed. The precipitate was suction filtrated and washed four times with 25 mL of 3% Na2CO3 solution to remove unreacted thioglycolic acid before being thoroughly dried to yield the products 6ae, 7a, and 7b.

4.1.1. 4-((2-Oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4-azaspiro[4.4]nonan-3-one (6a)

White ppt (69%), m.p 215–17 °C; 1H NMR (DMSO-d6): δH = 11.01 (bs, 1H; NH-1), 10.07 (bs, 1H; NH-4b), 7.97 (d, J = 8.1 Hz; 1H, H-5), 7.47 (dd, J = 8.1, 7.2 Hz; 1H, H-7), 7.28 (d, J = 7.8 Hz; 1H, H-8), 7.13 (t, J = 8, 7.2 Hz, H-6), 6.09 (s, 1H; H-3), 3.74 (s, 2H; H-2′), 1.1–1.95 ppm (m, 8H; cyclopentyl-CH2), 13C NMR (DMSO-d6): δC = 167.54 (C-3′), 162.89 (C-2), 149.69 (C-4), 139.05 (C-8a), 130.64 (C-7), 122.09 (C-5), 120.93 (C-6), 115.54 (C-8), 111.98 (C-4a), 92.64 (C-3), 54.40 (C-2′), 47.12 (C-5′), 34.11, 26.11, 25.25 ppm (Cyclopentyl-CH2), 15N NMR (DMSO-d6): δN = 142 (N-1), 128,6 (N-4b), 93.3 ppm (N-4′). m/z, Ms = 315 (M+, 12). Anl. Calcd. For C16H17N3O2S: C, 60.93; H, 5.43; N, 13.32; S, 10.17. Found: C, 61.08; H, 5.33; N, 13.43; S, 10.29.

4.1.2. 4-((2-Oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4-azaspiro[4.5]decan-3-one (6b)

White ppt (70%), m.p 194–96 °C; 1H NMR (DMSO-d6): δH = 11.05 (bs, 1H; NH-1), 10.09 (bs, 1H; NH-4b), 7.97 (m, 1H; H-5), 7.49 (m, 1H; H-6), 7.27 (m, 1H; H-7), 7.12 (m, 1H; H-8), 6.11 (s, 1H; H-3), 3.74 (s, 2H; H-2′), 1.11–1.92 ppm (m, 10H; cyclohexyl-CH2), 13C NMR (DMSO-d6): δC = 167.56 (C-3′), 162.85 (C-2), 149.76 (C-4), 139.06 (C-8a), 130.34 (C-7), 122.09 (C-5), 120.69 (C-6), 115.53 (C-8), 111.99 (C-4a), 92.66 (C-3), 54.42 (C-2′), 47.17 (C-5′), 34.18, 26.17 ppm (Cyclohexyl-CH), 15N NMR (DMSO-d6): δN = 142.1 (N-1), 131,7 (N-4b), 93.5 ppm (N-4′). m/z, Ms = 329 (M+, 10). Anl. Calcd. For C17H19N3O2S: C, 61.98; H, 5.81; N, 12.76; S, 9.73. Found: C, 62.11; H, 5.77; N, 12.83; S, 9.66.

4.1.3. 4-((6-Methyl-2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4-azaspiro[4.5]decan-3-one (6c)

White ppt (75%), m.p 160–62 °C; 1H NMR (DMSO-d6): δH = 10.93 (bs, 1H; NH-1), 10.05 (bs, 1H; NH-4b), 7.78(s, 1H, H-8), 7.33 (m, 1H; H-8), 7.17 (m, 1H; H-5), 6.07 (s, 1H; H-3), 3.24 (s, 2H; H-2′), 1.11–1.93 (m, 10H; Ali-CH), 2.11 ppm (s, 3H, CH3), 13C NMR (DMSO-d6): δC = 167.49 (C-3′), 162.15 (C-2), 149.75 (C-4), 139.16 (C-8a), 131.48 (C-7), 129.60 (C-6), 121.77 (C-5), 115.45 (C-8), 111.84 (C-4a), 92.61 (C-3), 54.33 (C-2′), 46.98 (C-5′), 33.90, 25.98 (Cyclohexyl-CH), 20.58 ppm (CH3). m/z, Ms = 343 (M+, 10). Anl. Calcd. For C18H21N3O2S: C, 62.95; H, 6.16; N, 12.23; S, 9.34. Found: C, 63.10; H, 6.22; N, 12.43; S, 9.21.

4.1.4. 4-((6-Methoxy-2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4-azaspiro[4.4]nonan-3-one (6d)

White ppt (67%), m.p 236–38 °C; 1H NMR (DMSO-d6): δH = 11.12 (bs, 1H; NH-1), 10.07 (bs, 1H; NH-4b), 7.99–7.11 (m, 3H; H-5,7,8), 6.06 (s, 1H; H-3), 3.70 (s, 3H; OMe), 3.35 (s, 2H; H-2′), 1.12–1.95 ppm (m, 8H; Cyclpentyl-CH2), 13C NMR (DMSO-d6): δC = 167.52 (C-3′), 162.84 (C-2), 149.11 (C-4), 136.07 (C-8a), 133.40 (C-6), 128.28 (C-7), 119.52 (C-5), 115.52 (C-8), 111.98 (C-4a), 92.67 (C-3), 55.55 (C-2′), 54.94 (OMe), 47.44 (C-5′), 32.77, 25.90, 23.77 ppm (Cyclopentyl-CH). Anl. Calcd. For C17H19N3O3S: C, 59.11; H, 5.54; N, 12.17; S, 9.28. Found: C, 59.23; H, 5.44; N, 11.99; S, 9.18.

4.1.5. 4-((6-Methoxy-2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4-azaspiro[4.5]decan-3-one (6e)

White ppt (79%), m.p 238–40 °C; 1H NMR (DMSO-d6): δH = 10.93 (bs, 1H; NH-1), 10.07 (bs, 1H; NH-4b), 7.50–7.34 (m, 1H; H-,8), 7.22–7.13 (m, 2H; H-5,7), 6.06 (s, 1H; H-3), 3.71 (s, 3H; OMe), 3.33 (s, 2H; H-2′), 1.14–1.92 ppm (m, 10H; Cyclohexyl-CH2), 13C NMR (DMSO-d6): δC = 167.60 (C-3′), 162.49 (C-2), 149.49 (C-4), 136.90 (C-8a), 133.46 (C-6), 128.28 (C-7), 119.52 (C-5), 116.82 (C-8), 112.29 (C-4a), 92.99 (C-3), 55.62 (C-2′), 55.42 (OMe), 47.56 (C-5′), 33.0, 25.94, 23.91 ppm (Cyclohexyl-CH). Anl. Calcd. For C18H21N3O3S: C, 60.15; H, 5.89; N, 11.69; S, 8.92. Found: C, 60.25; H, 5.77; N, 11.80; S, 8.88.

4.1.6. 7,7,9,9-Tetramethyl-4-((2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4,8-diazaspiro-[4.5]decan-3-one (7a)

White ppt (78%), m.p 165–67 °C; 1H NMR (DMSO-d6): δH = 11.07 (bs, 1H; NH-1), 10.26 (s, 1H; NH-4b), 7.78 (t, 1H; H-5), 7.65 (m,1H; H-7), 7.05 (m, 2H; NH, H-6), 6.05 (s, 1H; H-3), 3.51 (s, 2H; H-2′), 2.67 (s, 4; H-6′), 2.44 (s, 12H; 4CH3), 13C NMR (DMSO-d6): δC = 167.57 (C-3′), 162.68 (C-2), 149.68 (C-4), 139.02 (C-8a), 130.35 (C-7), 122.14 (C-5), 120.72 (C-6), 115.54 (C-8), 111.99 (C-4a), 92.87 (C-3), 55.55 (C-2′), 47.11 (C-5′),34.0 (C-6′), 26.33, 25.25 ppm (CH3). Anl. Calcd. For C20H26N4O2S: C, 62.15; H, 6.78; N, 14.50; S, 8.30. Found: C, 62.29; H, 6.88; N, 14.38; S, 8.36.

4.1.7. 8-Benzyl-4-((2-oxo-1,2-dihydroquinolin-4-yl)amino)-1-thia-4,8-diazaspiro[4.5]decan-3-one (7b)

White ppt (71%), m.p 180–82 °C; 1H NMR (DMSO-d6): δH = 11.01 (bs; 1H, NH-1), 10.09 (s; 1H, NH-4b), 7.98 (d, J = 8.1 Hz; 1H, H-5), 7.46 (dd, J = 7.7, 7.5 Hz; 1H, H-7), 7.35 (m; 4H, H-o, m), 7.28 (m; 3H, H-8, p), 7.12 (dd, J = 7.7, 7.4 Hz; 1H, H-6), 6.08 (s; 1H, H-3), 3.71 (s; 2H, H-2′), 3.56 (s; 2H, H-8″), 2.67 (t, J = 5.6 Hz; 2H, H-6′/6″), 2.55 (m; 4H, H-7′/7″), 2.44 ppm (t, J =5.4 Hz; 2H, H-6″/6′), 13C NMR (DMSO-d6): δC = 167.33 (C-3′), 162.90 (C-2), 149.51 (C-4), 139.23 (C-8a), 138.31 (C-i), 130.13 (C-7), 128.73 (C-o), 128.16 (C-m), 126.93 (C-p), 122.44 (C-5), 120.31 (C-6), 115.45 (C-8), 112.26 (C-4a), 92.33 (C-3), 61.36 (C-8″), 54.11 (C-2′), 53.23, 52.07 (C-7′/7″), 47.33 (C-5′), 34.46 (C-6′/6″), 27.26 ppm (C-6″/6′). Anl. Calcd. For C23H24N4O2S: C, 65.69; H, 5.75; N, 13.32; S, 7.62. Found: C, 65.51; H, 5.89; N, 13.22; S, 7.55.

4.2. Biology

4.2.1. Cell Viability Assay

The normal human mammary gland epithelial (MCF-10A) cell line was used to test the viability of the new compounds [46,47]. See Appendix SA (Supplementary File).

4.2.2. Antiproliferative Assay

The antiproliferative activity of 6ae, 7a, and 7b was tested against the four human cancer cell lines, Panc-1 (pancreatic cancer cell line), MCF-7 (breast cancer cell line), HT-29 (colon cancer cell line), and A-549 (lung cancer cell line), using the MTT assay and Erlotinib as the reference drug [48,49]. See Appendix SA (Supplementary File).

4.2.3. EGFR Inhibitory Assay

Compounds 6a–e, 7a, and 7b were tested for EGFR inhibitory activity as a potential target for their antiproliferative activity [50]. See Appendix SA (Supplementary File).

4.2.4. BRAFV600E Inhibitory Assay

The anti-BRAFV600E activity of compounds 6a–e, 7a, and 7b was also investigated in vitro [51] using Erlotinib as a reference compound. See Appendix SA (Supplementary File).

4.2.5. Apoptotic Markers Assay

Caspase 3 Activation Assay

Compounds 6b and 7b were tested as caspase-3 activators against the human pancreatic (Panc-1) cancer cell line [52,53,54]. See Appendix SA (Supplementary File).

Caspase-8, Bax and Bcl-2 Levels Assay

Compounds 6b and 7b were further investigated for their impact on the caspase-8, Bax, and Bacl-2 levels against the Panc-1 cancer cell line and staurosporine as a reference [55]. See Appendix SA (Supplementary File).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph16030467/s1.

Author Contributions

E.M.E.-S. and B.G.M.Y.: Conceptualization, writing, and editing, L.H.A.-W.: editing and revision; M.M.H.: writing the draft and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project Number (PNURSP2023R3), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data will be provided upon request.

Acknowledgments

The authors express their gratitude to Princess Nourah bint Abdulrahman University Researchers Supporting Project Number (PNURSP2023R3), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors reported no potential conflict of interest(s).

References

  1. Stanković, T.; Dinić, J.; Podolski-Renić, A.; Musso, L.; Burić, S.S.; Dallavalle, S.; Pešić, M. Dual Inhibitors as a New Challenge for Cancer Multidrug Resistance Treatment. Curr. Med. Chem. 2019, 26, 6074–6106. [Google Scholar] [CrossRef] [PubMed]
  2. Raghavendra, N.M.; Pingili, D.; Kadasi, S.; Mettu, A.; Prasad, S.V.U.M. Dual or multi-targeting inhibitors: The next generation anti-cancer agents. Eur. J. Med. Chem. 2018, 143, 1277–1300. [Google Scholar] [CrossRef] [PubMed]
  3. Hughes, D.; Andersson, D.I. Evolutionary consequences of drug resistance: Shared principles across diverse targets and organisms. Nat. Rev. Genet. 2015, 16, 459–471. [Google Scholar] [CrossRef] [PubMed]
  4. Staunton, J.E.; Jin, X.; Lee, M.S.; Zimmermann, G.R.; Borisy, A.A. Synergistic drug combinations tend to improve therapeutically relevant selectivity. Nat. Biotechnol. 2009, 27, 659–666. [Google Scholar] [CrossRef]
  5. Shah, K.N.; Bhatt, R.; Rotow, J.; Rohrberg, J.; Olivas, V.; Wang, V.E.; Hemmati, G.; Martins, M.M.; Maynard, A.; Kuhn, J.; et al. Aurora kinase A drives the evolution of resistance to third-generation EGFR inhibitors in lung cancer. Nat. Med. 2019, 25, 111–118. [Google Scholar] [CrossRef]
  6. Notarangelo, T.; Sisinni, L.; Condelli, V.; Landriscina, M. Dual EGFR and BRAF blockade overcomes resistance to vemurafenib in BRAF mutated thyroid carcinoma cells. Cancer Cell Int. 2017, 17, 86–94. [Google Scholar] [CrossRef] [Green Version]
  7. Mondaca, S.; Lacouture, M.; Hersch, J.; Yaeger, R. Balancing RAF, MEK, and EGFR inhibitor doses to achieve clinical responses and modulate toxicity in BRAF V600E colorectal cancer. JCO Precis. Oncol. 2018, 10, 1–5. [Google Scholar] [CrossRef]
  8. Zhang, Q.; Diao, Y.; Wang, F.; Fu, Y.; Tang, F.; You, Q.; Zhou, H. Design and discovery of 4-anilinoquinazoline ureas as multikinase inhibitors targeting BRAF, VEGFR-2 and EGFR. MedChemComm 2013, 4, 979–986. [Google Scholar] [CrossRef]
  9. Okaniwa, M.; Hirose, M.; Imada, T.; Ohashi, T.; Hayashi, Y.; Miyazaki, T.; Arita, T.; Yabuki, M.; Kakoi, K.; Kato, J.; et al. Design and synthesis of novel DFG-out RAF/vascular endothelial growth factor receptor 2 (VEGFR2) inhibitors. 1. Exploration of [5,6]-fused bicyclic scaffolds. J. Med. Chem. 2012, 55, 3452–3478. [Google Scholar] [CrossRef]
  10. Roskoski, R., Jr. Properties of FDA-approved small molecule protein kinase inhibitors. Pharmacol. Res. 2019, 144, 19–50. [Google Scholar] [CrossRef]
  11. Prajapati, S.M.; Patel, K.D.; Vekariya, R.H.; Panchal, S.N.; Patel, H.D. Recent advances in the synthesis of quinolines: A review. RSC Adv. 2014, 4, 24463–24476. [Google Scholar] [CrossRef]
  12. Navneetha, O.; Deepthi, K.; Rao, A.M.; Jyostna, T.S. A review on chemotherapeutic activities of quinolone. Int. J. Pharm. Chem. Biol. Sci. 2017, 7, 364–372. [Google Scholar]
  13. Barlési, F.; Tchouhadjian, C.; Doddoli, C.; Villani, P.; Greillier, L.; Kleisbauer, J.P.; Thomas, P.; Astoul, P. Gefitinib (ZD1839, Iressa®) in non-small-cell lung cancer: A review of clinical trials from a daily practice perspective. Fund. Clin. Pharmacol. 2005, 19, 385–393. [Google Scholar] [CrossRef]
  14. Iyer, R.; Bharthuar, A. A review of erlotinib–an oral, selective epidermal growth factor receptor tyrosine kinase inhibitor. Exp. Opin. Pharmacother. 2010, 11, 311–320. [Google Scholar] [CrossRef]
  15. Bao, B.; Mitrea, C.; Wijesinghe, P.; Marchetti, L.; Girsch, E.; Farr, R.L.; Boerner, J.L.; Mohammad, R.; Dyson, G.; Terlecky, S.R.; et al. Treating triple negative breast cancer cells with erlotinib plus a select antioxidant overcomes drug resistance by targeting cancer cell heterogeneity. Sci. Rep. 2017, 7, 44125. [Google Scholar] [CrossRef] [Green Version]
  16. 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]
  17. Wissner, A.; Berger, D.M.; Boschelli, D.H.; Floyd, M.B.; Greenberger, L.M.; Gruber, B.C.; Johnson, B.D.; Mamuya, N.; Nilakantan, R.; Reich, M.F.; et al. 4-Anilino-6,7-dialkoxyquinoline-3-carbonitrile inhibitors of epidermal growth factor receptor kinase and their bioisosteric relationship to the 4-anilino-6,7-dialkoxyquinazoline inhibitors. J. Med. Chem. 2000, 43, 3244–3256. [Google Scholar] [CrossRef]
  18. Wissner, A.; Mansour, T.S. The development of HKI-272 and related compounds for the treatment of cancer. Arch. Pharm. 2008, 341, 465–477. [Google Scholar] [CrossRef]
  19. Kiesel, B.F.; Parise, R.A.; Wong, A.; Keyvanjah, K.; Jacobs, S.; Beumer, J.H. LC–MS/MS assay for the quantitation of the tyrosine kinase inhibitor neratinib in human plasma. J. Pharm. Biomed. Anal. 2017, 134, 130–136. [Google Scholar] [CrossRef] [Green Version]
  20. Pisaneschi, F.; Nguyen, Q.-D.; Shamsaei, E.; Glaser, M.; Robins, E.; Kaliszczak, M.; Smith, G.; Spivey, A.C.; Aboagye, E.O. Development of a new epidermal growth factor receptor positron emission tomography imaging agent based on the 3-cyanoquinoline core: Synthesis and biological evaluation. Bioorg. Med. Chem. 2010, 18, 6634–6645. [Google Scholar] [CrossRef]
  21. Lü, S.; Zheng, W.; Ji, L.; Luo, Q.; Hao, X.; Li, X.; Wang, F. Synthesis, characterization, screening and docking analysis of 4-anilinoquinazoline derivatives as tyrosine kinase inhibitors. Eur. J. Med. Chem. 2013, 61, 84–94. [Google Scholar] [CrossRef]
  22. Luethi, D.; Durmus, S.; Schinkel, A.H.; Schellens, J.H.M.; Beijnen, J.H.; Sparidans, R.W. Liquid chromatography–tandem mass spectrometry assay for the EGFR inhibitor pelitinib in plasma. J. Chromatogr. B 2013, 934, 22–25. [Google Scholar] [CrossRef] [PubMed]
  23. Pawar, V.G.; Sos, M.L.; Rode, H.B.; Rabiller, M.; Heynck, S.; van Otterlo, W.A.L.; Thomas, R.K.; Rauh, D. Synthesis and biological evaluation of 4-anilinoquinolines as potent inhibitors of epidermal growth factor receptor. J. Med. Chem. 2010, 53, 2892–2901. [Google Scholar] [CrossRef] [PubMed]
  24. Elbastawesy, M.A.I.; Aly, A.A.; Ramadan, M.; Elshaier, Y.A.M.M.; Youssif, B.G.M.; Brown, A.B.; Abuo-Rahma, G.E.A. Novel Pyrazoloquinolin-2-ones: Design, Synthesis, Docking Studies, and Biological Evaluation as Antiproliferative EGFR- TK Inhibitors. Bioorg. Chem. 2019, 90, 103045–103060. [Google Scholar] [CrossRef] [PubMed]
  25. Mohassab, A.M.; Hassan, H.A.; Abdelhamid, D.; Gouda, A.M.; Youssif, B.G.M.; Tateishi, H.; Fujita, M.; Otsuka, M.; Abdel-Aziz, M. Design and Synthesis of Novel quinoline/chalcone/1,2,4-triazole hybrids as potent antiproliferative agent targeting EGFR and BRAFV600E kinases. Bioorg. Chem. 2021, 106, 104510. [Google Scholar] [CrossRef]
  26. Zheng, Y.; Tice, C.M.; Singh, S.B. The use of spirocyclic scaffolds in drug discovery. Bioorg. Med. Chem. Lett. 2014, 24, 3673–3682. [Google Scholar] [CrossRef] [Green Version]
  27. Batista, V.F.; Pinto, D.C.G.A.; Silva, A.M.S. Recent in vivo advances of spirocyclic scaffolds for drug discovery. Expert Opin. Drug Discov. 2022, 17, 603–618. [Google Scholar] [CrossRef]
  28. Hiesinger, K.; Dar’in, D.; Proschak, E.; Krasavin, M. Spirocyclic Scaffolds in Medicinal Chemistry. J. Med. Chem. 2021, 64, 150–183. [Google Scholar] [CrossRef]
  29. Ayati, A.; Emami, S.; Asadipour, A.; Shafiee, A.; Foroumadi, A. Recent applications of 1,3-thiazole core structure in the identification of new lead compounds and drug discovery. Eur. J. Med. Chem. 2015, 97, 699–718. [Google Scholar] [CrossRef]
  30. Sharma, P.C.; Jain, A.; Yar, M.S.; Pahwa, R.; Singh, J.; Chanalia, P. Novel fluoroquinolone derivatives bearing N-thiomide linkage with 6-substituted-2-aminobenzothiazoles: Synthesis and antibacterial evaluation. Arab. J. Chem. 2017, 10, S568–S575. [Google Scholar] [CrossRef] [Green Version]
  31. Petrou, A.; Fesatidou, M.; Geronikaki, A. Thiazole Ring—A Biologically Active Scaffold. Molecules 2021, 26, 3166. [Google Scholar] [CrossRef]
  32. Othman, I.M.M.; Alamshany, Z.M.; Tashkandi, N.Y.; Gad-Elkareem, M.A.M.; Abd El-Karim, S.S.; Nossier, E.S. Synthesis and biological evaluation of new derivatives of thieno-thiazole and dihydrothiazolo-thiazole scaffolds integrated with a pyrazoline nucleus as anticancer and multi-targeting kinase inhibitor. RSC Adv. 2022, 12, 561–577. [Google Scholar] [CrossRef]
  33. Nafie, M.S.; Kishk, S.M.; Mahgoub, S.; Amer, A.M. Quinoline-based thiazolidinone derivatives as potent cytotoxic and apoptosis-inducing agents through EGFR inhibition. Chem. Biol. Drug Des. 2022, 99, 547–560. [Google Scholar] [CrossRef]
  34. Xiong, L.; He, H.; Fan, M.; Hu, L.; Wang, F.; Song, X.; Shi, S.; Qi, B. Discovery of novel conjugates of quinoline and thiazolidinone urea as potential anti-colorectal cancer agent. J. Enzym. Inhib. Med. Chem. 2022, 37, 2334–2347. [Google Scholar] [CrossRef]
  35. Yadav, J.; Chaudhary, R.P. A review on advances in synthetic methodology and biological profile of spirothiazolidin-4-ones. J. Heterocycl. Chem. 2022, 59, 1839–1878. [Google Scholar] [CrossRef]
  36. Lozynskyi, A.; Zimenkovsky, B.; Lesyk, R. Synthesis and Anticancer Activity of New Thiopyrano[2,3-d]thiazoles Based on Cinnamic Acid Amides. Sci. Pharm. 2014, 82, 723–733. [Google Scholar] [CrossRef] [Green Version]
  37. Hu-Lieskovan, S.; Mok, S.; Homet Moreno, B.; Tsoi, J.; Robert, L.; Goedert, L.; Pinheiro, E.; Koya, R.; Graeber, T.; Comin-Anduix, B.; et al. Improved antitumor activity of immunotherapy with BRAF and MEK inhibitors in BRAFV600E melanoma. Sci. Transl. Med. 2015, 7, 279ra41. [Google Scholar] [CrossRef] [Green Version]
  38. Abdel-Maksoud, M.S.; Kim, M.-R.; El-Gamal, M.I.; Gamal El-Din, M.M.; Tae, J.; Choi, H.S.; Lee, K.-T.; Yoo, K.H.; Oh, C.-H. Design, synthesis, in vitro antiproliferative evaluation, and kinase inhibitory effects of a new series of imidazo[2,1-b]thiazole derivatives. Eur. J. Med. Chem. 2015, 95, 453–463. [Google Scholar] [CrossRef]
  39. Aly, A.A.; Alshammari, M.B.; Ahmad, A.; Gomaa, H.A.M.; Youssif, B.G.M.; Bräse, S.; Ibrahim, M.A.A.; Mohamed, A.H. Design, synthesis, docking, and mechanistic studies of new thiazolyl/thiazolidinylpyrimidine-2,4-dione antiproliferative agents. Arab. J. Chem. 2023, 16, 104612. [Google Scholar] [CrossRef]
  40. Al-Wahaibi, L.H.; Gouda, A.M.; Abou-Ghadir, O.F.; Salem, O.I.A.; Ali, A.T.; Farghaly, H.S.; Abdelrahman, M.H.; Trembleau, L.; Abdu-Allah, H.H.M.; Youssif, B.G.M. Design and synthesis of novel 2,3-dihydropyrazino[1,2-a]indole-1,4-dione derivatives as antiproliferative EGFR and BRAFV600E dual inhibitors. Bioorg. Chem. 2020, 104, 104260. [Google Scholar] [CrossRef]
  41. Gomaa, H.A.M.; Shaker, M.E.; Alzarea, S.I.; Hendawy, O.M.; Mohamed, F.A.M.; Gouda, A.M.; Ali, A.T.; Morcoss, M.M.; Abdelrahman, M.H.; Trembleau, L.; et al. Optimization and SAR investigation of novel 2,3-dihydropyrazino[1,2-a]indole-1,4-dione derivatives as EGFR and BRAFV600E dual inhibitors with potent antiproliferative and antioxidant activities. Bioorg. Chem. 2022, 120, 105616. [Google Scholar] [CrossRef] [PubMed]
  42. Alshammari, M.B.; Aly, A.A.; Youssif, B.G.M.; Bräse, S.; Ahmad, A.; Brown, A.B.; Ibrahim, M.A.A.; Mohamed, A.H. Design and synthesis of new thiazolidinone/uracil derivatives as antiproliferative agents targeting EGFR and/or BRAFV600E. Front. Chem. 2022, 10, 1076383. [Google Scholar] [CrossRef] [PubMed]
  43. Buckle, D.R.; Cantello, B.C.C.; Smith, H.; Spicer, B.A. 4-Hydroxy-3-nitro-2-quinolones and related compounds as inhibitors of allergic reactions. J. Med. Chem. 1975, 18, 726–732. [Google Scholar] [CrossRef] [PubMed]
  44. Bhudevi, B.; Ramana, P.V.; Mudiraj, A.; Reddy, A.R. Synthesis of 4-hydroxy-3-formylideneamino-lH/methyl/phenylquinolin2-ones. Indian J. Chem. B 2009, 48, 255–260. [Google Scholar]
  45. Yang, Z.; Sun, P. Compare of three ways of synthesis of simple Schiff bas. Molbank 2006, 2006, M514. [Google Scholar] [CrossRef] [Green Version]
  46. Mahmoud, M.A.; Mohammed, A.F.; Salem, O.I.A.; Gomaa, H.A.M.; Youssif, B.G.M. New 1,3,4-oxadiazoles linked 1,2,3-triazole moiety as antiproliferative agents targeting EGFR-TK. Arch. Der Pharm. 2022, 355, e2200009. [Google Scholar] [CrossRef]
  47. Ramadan, M.; Abd El-Aziz, M.; Elshaier, Y.A.M.M.; Youssif, B.G.M.; Brown, A.B.; Fathy, H.M.; Aly, A.A. Design and synthesis of new pyranoquinolinone heteroannulated to triazolopyrimidine of potential apoptotic antiproliferative activity. Bioorg. Chem. 2020, 105, 104392. [Google Scholar] [CrossRef]
  48. Al-Sanea, M.M.; Gotina, L.; Mohamed, M.F.A.; Parambi, D.G.T.; Gomaa, H.A.M.; Mathew, B.; Youssif, B.G.M.; Alharbi, K.S.; Elsayed, Z.M.; Abdelgawad, M.A.; et al. Design, Synthesis and Biological Evaluation of New HDAC1 and HDAC2 Inhibitors Endowed with Ligustrazine as a Novel Cap Moiety. Drug Des. Dev. Ther. 2020, 14, 497–508. [Google Scholar] [CrossRef] [Green Version]
  49. Hisham, M.; Hassan, H.A.; Gomaa, H.A.M.; Youssif, B.G.M.; Hayallah, A.M.; Abdel-Aziz, M. Structure-based design, synthesis and antiproliferative action of new quinazoline-4-one/chalcone hybrids as EGFR inhibitors. J. Mol. Struct. 2022, 1254, 132422. [Google Scholar] [CrossRef]
  50. Mohamed, F.A.M.; Gomaa, H.A.M.; Hendawy, O.M.; Ali, A.T.; Farghaly, H.S.; Gouda, A.M.; Abdelazeem, A.H.; Abdelrahman, M.H.; Trembleau, L.; Youssif, B.G.M. Design, synthesis, and biological evaluation of novel EGFR inhibitors containing 5-chloro-3-hydroxymethyl-indole-2-carboxamide scaffold with apoptotic antiproliferative activity. Bioorg. Chem. 2021, 112, 104960. [Google Scholar] [CrossRef]
  51. Abou-Zied, H.A.; Beshr, E.A.M.; Gomaa, H.A.M.; Mostafa, Y.A.; Youssif, B.G.M.; Hayallah, A.M.; Abdel-Aziz, M. Discovery of new cyanopyridine/chalcone hybrids as dual inhibitors of EGFR/BRAFV600E with promising antiproliferative properties. Arch. Der Pharm. 2022, 355, e2200464. [Google Scholar] [CrossRef]
  52. Hisham, M.; Youssif, B.G.M.; Osman, E.E.A.; Hayallah, A.M.; Abdel-Aziz, M. Synthesis and biological evaluation of novel xanthine derivatives as potential apoptotic antitumor agents. Eur. J. Med. Chem. 2019, 176, 117–128. [Google Scholar] [CrossRef] [PubMed]
  53. Martin, S.J. Caspases: Executioners of apoptosis. Pathobiol. Hum. Dis. 2014, 145, 145–152. [Google Scholar]
  54. Abdelbaset, M.S.; Abdel-Aziz, M.; Abuo-Rahma, G.E.A.; Abdelrahman, M.H.; Ramadan, M.; Youssif, B.G.M. Novel quinoline derivatives carrying nitrones/oximes nitric oxide donors: Design, synthesis, antiproliferative and caspase-3 activation activities. Arch. Der Pharm. 2018, 352, 1800270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Abou-Zied, H.A.; Youssif, B.G.M.; Mohamed, M.F.A.; Hayallah, A.M.; Abdel-Aziz, M. EGFR inhibitors and apoptotic inducers: Design, synthesis, anticancer activity and docking studies of novel xanthine derivatives carrying chalcone moiety as hybrid molecules. Bioorg. Chem. 2019, 89, 102997. [Google Scholar] [CrossRef]
Figure 1. Structure of Erlotinib, Gefitinib, Neratinib, Pelitinib, and compounds IIII.
Figure 1. Structure of Erlotinib, Gefitinib, Neratinib, Pelitinib, and compounds IIII.
Pharmaceuticals 16 00467 g001
Figure 2. Structures of some thiazole-based anticancer agents IVVI.
Figure 2. Structures of some thiazole-based anticancer agents IVVI.
Pharmaceuticals 16 00467 g002
Figure 3. Structures of new compounds 6ae, 7a, and 7b.
Figure 3. Structures of new compounds 6ae, 7a, and 7b.
Pharmaceuticals 16 00467 g003
Scheme 1. Synthesis of azaspiroalkan-3-one 6ae.
Scheme 1. Synthesis of azaspiroalkan-3-one 6ae.
Pharmaceuticals 16 00467 sch001
Figure 4. Structure of compound 6b and its isomeric structure 6b′.
Figure 4. Structure of compound 6b and its isomeric structure 6b′.
Pharmaceuticals 16 00467 g004
Figure 5. 1H -15N HMBC for compound 6b.
Figure 5. 1H -15N HMBC for compound 6b.
Pharmaceuticals 16 00467 g005
Scheme 2. Synthesis of azaspiroalkan-3-one 7a,b.
Scheme 2. Synthesis of azaspiroalkan-3-one 7a,b.
Pharmaceuticals 16 00467 sch002
Figure 6. Structure of compound 7b.
Figure 6. Structure of compound 7b.
Pharmaceuticals 16 00467 g006
Scheme 3. Suggested mechanism which rationalizes compounds 6ae.
Scheme 3. Suggested mechanism which rationalizes compounds 6ae.
Pharmaceuticals 16 00467 sch003
Table 1. Spectral data for compound 7b.
Table 1. Spectral data for compound 7b.
1H NMR1H-1H COSYAssignment
11.01 (bs; 1H)6.08NH-1
10.09 (s; 1H) NH-4b
7.98 (d, J = 8.1; 1H)7.46, 7.12H-5
7.46 (dd, J = 7.7, 7.5; 1H)7.98, 7.28, 7.12H-7
7.35 (m; 4H)7.28H-o, m
7.28 (m; 3H)7.46, 7.35, 7.12H-8, p
7.12 (dd, J = 7.7, 7.4; 1H)7.98, 7.46, 7.28H-6
6.08 (s; 1H)11.01H-3
3.71 (s; 2H) H-2′
3.56 (s; 2H) H-8″
2.67 (t, J = 5.6; 2H)2.55, 2.44H-6′/6″
2.55 (m; 4H)2.67, 2.44H-7′,7″
2.44 (t, J = 5.4; 2H)2.67, 2.55H-6″/6′
13C NMRHSQCHMBCAssignment
167.33 10.09, 3.71C-3′
162.90 10.09, 6.08C-2
149.51 10.09, 7.98, 7.28, 6.08C-4
139.23 7.98, 7.46C-8a
138.31 7.35, 7.28, 3.56C-i
130.137.467.98, 7.46, 7.28C-7
128.737.357.35, 7.35, 3.56C-o
128.167.357.35, 7.35C-m
126.937.287.28C-p
122.447.987.98, 7.46, 6.08C-5
120.317.127.28, 7.12C-6
115.457.287.46, 7.12C-8
112.26 10.09, 7.98, 7.28, 7.12, 6.08C-4a
92.336.0810.09, 6.08C-3
61.36 3.56, 2.55C-8″
54.113.712.67, 244C-2′
53.23, 52.072.553.56, 2.67, 2.55, 2.55, 2.44C-7′,7″
47.333.562.67, 2.44C-5′
34.462.442.67, 2.55C-6′/6″
27.262.672.67, 2.55, 2.55C-6″/6′
15N NMR:HSQCHMBCAssignment
141.911.007.28, 6.08N-1
132.010.0910.09, 6.08N-4b
49.7 2.67, 2.44N-8′
Table 2. IC50 of compounds 6ae, 7a, 7b, and Erlotinib against four cancer cell lines.
Table 2. IC50 of compounds 6ae, 7a, 7b, and Erlotinib against four cancer cell lines.
Compd.Cell Viability %Antiproliferative Activity IC50 ± SEM (µM)
A-549MCF-7Panc-1HT-29Average
(GI50)
6a9040 ± 444 ± 446 ± 448 ± 445
6b8933 ± 335 ± 336 ± 336 ± 335
6c9172 ± 774 ± 773 ± 772 ± 773
6d9148 ± 452 ± 554 ± 554 ± 552
6e9236 ± 340 ± 342 ± 442 ± 440
7a8980 ± 882 ± 881 ± 881 ± 881
7b9130 ± 334 ± 332 ± 332 ± 332
ErlotinibND30 ± 340 ± 330 ± 330 ± 333
ND: Not Determined.
Table 3. IC50 of compounds 6ae, 7a, 7b, and Erlotinib against EGFR and BRAFV600E.
Table 3. IC50 of compounds 6ae, 7a, 7b, and Erlotinib against EGFR and BRAFV600E.
Compd.EGFR Inhibition
IC50 ± SEM (nM)
BRAFV600E Inhibition
IC50 ± SEM (nM)
6a97 ± 07137 ± 12
6b84 ± 06108 ± 09
6c135 ± 11164 ± 15
6d104 ± 08159 ± 14
6e92 ± 07129 ± 10
7a149 ± 12187 ± 16
7b78 ± 0596 ± 8
Erlotinib80 ± 0560 ± 05
Table 4. Caspase-3 induction of compounds 6b and 7b.
Table 4. Caspase-3 induction of compounds 6b and 7b.
Compound NumberCaspase-3
Conc (Pg/mL)Fold Change
6b487.50 ± 47.50
7b544.50 ± 58.5
Staurosporine503.00 ± 48.0
Control65.501
Table 5. Caspase-8, Bax and Bcl-2 levels for compounds 6b, 7b, and Staurosporine.
Table 5. Caspase-8, Bax and Bcl-2 levels for compounds 6b, 7b, and Staurosporine.
Compound NumberCaspase-8BaxBcl-2
Conc (ng/mL)Fold ChangeConc (Pg/mL)Fold ChangeConc (ng/mL)Fold Reduction
6b1.5017220281.304
7b1.9021295371.005
Staurosporine1.8020280351.105
Control0.0918151
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

Al-Wahaibi, L.H.; El-Sheref, E.M.; Hammouda, M.M.; Youssif, B.G.M. One-Pot Synthesis of 1-Thia-4-azaspiro[4.4/5]alkan-3-ones via Schiff Base: Design, Synthesis, and Apoptotic Antiproliferative Properties of Dual EGFR/BRAFV600E Inhibitors. Pharmaceuticals 2023, 16, 467. https://doi.org/10.3390/ph16030467

AMA Style

Al-Wahaibi LH, El-Sheref EM, Hammouda MM, Youssif BGM. One-Pot Synthesis of 1-Thia-4-azaspiro[4.4/5]alkan-3-ones via Schiff Base: Design, Synthesis, and Apoptotic Antiproliferative Properties of Dual EGFR/BRAFV600E Inhibitors. Pharmaceuticals. 2023; 16(3):467. https://doi.org/10.3390/ph16030467

Chicago/Turabian Style

Al-Wahaibi, Lamya H., Essmat M. El-Sheref, Mohamed M. Hammouda, and Bahaa G. M. Youssif. 2023. "One-Pot Synthesis of 1-Thia-4-azaspiro[4.4/5]alkan-3-ones via Schiff Base: Design, Synthesis, and Apoptotic Antiproliferative Properties of Dual EGFR/BRAFV600E Inhibitors" Pharmaceuticals 16, no. 3: 467. https://doi.org/10.3390/ph16030467

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