Next Article in Journal
Failed Repurposing of Lysosomotropic Drugs for COVID-19 Treatment or Prevention
Previous Article in Journal
Drugs and Drug Candidates: A Challenging Project
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Selective Cytotoxic Activity against Human Breast Cancer MCF7 Cell Line and Molecular Docking of Some Chalcone-Dihydropyrimidone Hybrids

by
Eduardo B. Mass
1,
Carolina A. de Lima
2,
Marcelo G. M. D’Oca
3,
Juliana M. Sciani
4,
Giovanna B. Longato
2,† and
Dennis Russowsky
1,*,†
1
Laboratório de Sínteses Orgânicas, Instituto de Química, Universidade Federal do Rio Grande do Sul, Av. Bento Gonçalves 9500, Porto Alegre CEP 91501-970, RS, Brazil
2
Laboratório de Pesquisa em Farmacologia Molecular e Compostos Bioativos–Prédio VII, Universidade São Francisco, Avenida São Francisco de Assis 218, Bragança Paulista CEP 12916-900, SP, Brazil
3
Departamento de Química, Universidade Federal do Paraná, Centro Politécnico, Av. Coronel Francisco H. Santos 100, Curitiba CEP 81531-980, PR, Brazil
4
Laboratório Multidisciplinar de Pesquisas–Prédio V, Universidade São Francisco, Av. São Francisco de Assis 218, Bragança Paulista CEP 12916-900, SP, Brazil
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Drugs Drug Candidates 2022, 1(1), 3-21; https://doi.org/10.3390/ddc1010002
Submission received: 10 October 2022 / Revised: 17 November 2022 / Accepted: 23 November 2022 / Published: 1 December 2022
(This article belongs to the Section Medicinal Chemistry and Preliminary Screening)

Abstract

:
Designed Chalcone-Dihydropyrimidinone hybrid compounds were synthesized expeditiously. The hybridization was performed through the Copper-catalyzed Alkyne-Azide Cycloaddition (CuAAC) from the propargyloxy chalcones and azido-dihydropyrimidinones. The hybrid products were prepared in five steps with a 30–48% overall yield. Most of the compounds showed selective cytotoxicity and lower IC50 values (<10 µM) against MCF-7 (breast adenocarcinoma) cancer. Cytotoxicity was also observed against OVCAR-3 (ovary, adenocarcinoma), NCI/ADR-RES (ovary, multidrug-resistant adenocarcinoma), and U-251 (brain, glioblastoma) cell lines. The potency of the most active hybrids 9d, 9g, and 9h was greater than the individual parental compounds, suggesting the effectiveness of molecular hybridization on the cytotoxicity. Compounds 9d, 9g, and especially 9h showed high selectivity for breast cancer cells (MCF-7) regarding human keratinocytes (HaCaT). Molecular docking calculations for the 9d, 9g, and 9h hybrids in the active site of estrogen supported the hypothesis that the compounds act as ER-α antagonists, disrupting the cell proliferation process of MCF-7, corroborating the potency and selectivity observed for this tumoral cell line.

1. Introduction

Molecular hybridization is a strategy that consists of the combination of two or more pharmacophores through a covalent bond [1]. The hybrid molecules can affect different factors of the same disease, counterbalancing side effects, interacting with multiple targets (multifunctional), or amplifying the potency [2,3]. This approach has increasingly become a paradigm for the pharmaceutical industry not only in terms of financial costs but also due to the reduction of side effects caused by unwanted drug interactions in drug cocktails [4,5,6].
Many successful examples of highly bioactive compounds can be found in the literature due to the hybridization protocol of different pharmacophores, including natural-based hybrid molecules [7,8]. Remarkably antiproliferative activities [9], leishmanicidal [10], or inhibitory activity against acetylcholinesterase enzyme [11] are some of the prominent reported cases.
One interesting scaffold to explore in the design of hybrid compounds is the chalcone moiety. Chalcones are a group of plant-derived polyphenolic compounds belonging to the flavonoids family. The structural variety and plethora of biological activities of chalcones were exhaustively discussed in recent reviews [12,13,14]. Additionally, they are known to express high anticancer activity, notably as inhibitors of tubulin polymerization by binding in the colchicine site [15,16]. Over the last few years, intense research on chalcone hybrid compounds has been reported. New biologically active chalcone hybrids were synthesized by connecting different counterparts such as coumarins [17], sulfonamides [18], anthraquinones [19], indolizines [20], quinazolines [21,22], cinchonas [23], quinolines [24], pyrazolines [25] and benzimidazoles [26] among others. Despite the availability of different methods for preparing chalcones, they can be easily synthesized directly via base-catalyzed Claisen–Schmidt condensation from acetophenones and benzaldehydes [27]. Among the available methods to prepare the chalcone hybrids [28,29], the CuAAC (Copper Catalyzed Alkyne-Azide Cycloaddition) strategy [30,31] is an easy way to link the chalcone part to another molecular unity. The CuAAC methodology produces a 1,2,3-triazole heterocycle as the linker between the two individual parts as shown in the hybrids Chalcone-Benzyl (I) [32] Chalcone β-Lactam (II) [33] and Chalcone-Quinoline (III) [34] respectively (Figure 1).
Other privileged scaffolds in Medicinal Chemistry are the dihydropyrimidin-2-(thi)ones (DHPMs) also due to various biological activities they present. Monastrol (IV), a dihydropyrimidin-2-thione discovered by Mayer in early 1999s [35], became a relevant non-natural small molecule able to disturb the mitosis process by inhibiting the mitotic kinesin Eg5, a motor protein responsible for the bipolar spindle formation without affecting the tubulin dynamics [36]. Several recent reviews covered the synthesis and biological activities [37,38,39].
The DHPMs have been used as a template for developing diverse new anticancer hybrid molecules. The hybrids Monastrol-Fatty acids (V, cytotoxic against glioblastomas) [40] and DHPM-perillyl alcohol (VI, cytotoxic against OVCAR-3, ovary cell line) were investigated by our group [9]. More recently, a fluorescent dihydropyrimin-2-thinone-arylbenzoxazole hybrid (VII) was identified as promising against prostate cells (PC3 cell line). Due to the strong fluorescence of this compound, it was possible to observe their accumulation in the endoplasmic reticulum and the nuclei of cancer cells (Figure 2) [41].
Inspired by these observations, we propose a new design for hybrid molecules based on the combination of chalcones and DHPMs, in our quest for potential antitumoral activity. For this purpose, we installed an alkyne group in the chalcone component and an azide group in the DHPM. The CuAAC protocol was chosen to connect both pharmacophores via a triazole linkage. The unique properties of the triazole linker bring stability to the hybrid molecule adding the advantage of being a bioisostere, notably as an amide function [42].

2. Results and Discussion

2.1. Syntehsis of Chalcone-DHPM Hybrids

The propargyloxy chalcones were prepared according to the synthetic route shown in Scheme 1. The hydroxy benzaldehydes 1ae were alkylated with propargyl bromide to afford derivatives 2ae in 82–95% yields [43]. The aldehydes 2ae were further condensed with acetophenone (3) to give access to chalcones 4ae in 67–82% yield (Scheme 1) [44].
In parallel, the chloro-DHPMs 7a,b were prepared via multicomponent Biginelli reaction from benzaldehydes 1fg, ethyl 4-chloroacetoacetate (5), and urea (6) [45]. The Biginelli compounds 7ac were obtained in 67–72% yield. Next, they were converted to azido-DHPMs 8a,b in excellent yields, 82–91% in the reaction with NaN3 (Scheme 2) [46].
After that, the chalcones 4ae and DHPM 8a,b were hybridized through the CuAAC methodology in the presence of CuSO4·5H2O/sodium ascorbate system to generate in situ the active Cu(I) specie for the alkyne-azide 1,3-dipolar cycloaddition reaction [47]. The combination of each chalcone with all DHPMs afforded a small library of ten Chalcone-DHPM hybrids 9aj (Scheme 3). After purification by column chromatography, the hybrids were obtained in reasonable-to-good yields. All hybrids were fully characterized by conventional spectroscopic methods, including HRMS (High-Resolution Mass Spectrometry) for novel compounds.
The appearance of a singlet signal around 8.39–9.35 ppm in the 1H NMR spectra, assigned as the hydrogen of the triazole ring, confirms its formation as a linker. The results are shown in Table 1 below.

2.2. Cytotoxicity of Chalcone-DHPM Hybrids

The cytotoxicity of hybrids 9aj was evaluated against a panel of human cancer cell lines through the MTT assay [48]. Doxorubicin (DOXO) was used as the positive control. The results are presented in Table 2 below.
As shown in Table 2, breast adenocarcinoma (MCF-7) was the most sensitive cell line to hybrid compounds treatment, and IC50 values were very low, ranging from 4.7 to 14.6 µM. Besides being potent for MCF-7, compounds 9f and 9g were the most potent (<15 µM) for glioblastoma (U-251), and compounds 9a, 9b, and 9f presented low IC50 values (<15 µM) for ovarian adenocarcinoma (OVCAR-3). Compounds 9a and 9b presented lower IC50 values (6.4 and 14.5 µM, respectively) than the positive control doxorubicin (22.6 µM) for NCI/ADR-RES. This result is very relevant because this is an ovarian adenocarcinoma cell line that expresses the phenotype of resistance to multiple drugs, characterized by the high expression of glycoprotein-P (P-gp), an energy-dependent transmembrane efflux pump responsible for modulating multidrug resistance (MDR) [49]. Further studies with these molecules must be conducted on this cell line to investigate their role in MDR reversion.
The IC50 values of all hybrid compounds against non-tumoral cell line HaCaT were higher than the doxorubicin (IC50 = 1.6 µM). Furthermore, compound 9h presented a great cytotoxic effect against MCF-7 cells (IC50 = 5.8 µM) with no toxic effects for HaCaT (IC50 > 100 µM). The IC50 value of hybrid 9h against HaCaT is 62.9 times greater than the one found for doxorubicin. It might indicate a preference for this compound for tumoral cell lines and less toxicity against healthy cells. Considering that compounds 9d and 9g presented the lowest IC50 values for the most sensitive cell line MCF-7 and that compound 9h was not cytotoxic for HaCaT, these molecules were chosen for a comparative cytotoxicity profile study with their parental molecules.

2.3. Comparative Cytotoxicity Profile of Hybrids and Their Parental Molecules

In order to investigate the effect of molecular hybridization on the cytotoxicity, the chalcones 4bd, and DHPMs 8a,b (as constituent parts of the chosen molecules 9d, 9g, and 9h) were screened separately against U-251, MCF-7, and NCI/ADR-RES tumoral cell lines, as well as HaCaT.
As can be seen in Table 3, hybrids 9d and 9g were more cytotoxic than their parental chalcone and DHPM for all the cell lines evaluated, including the non-tumoral HaCaT. Hybridization turned molecules very potent for breast adenocarcinoma MCF-7 cell line (IC50 value of 4.72 and 4.89 µM) differing from their parental chalcone 4d e 4b and DHPM 8a e 8b, which showed the following IC50 values: >342, 80.33, >332 and >255 µM, respectively. Although the IC50 of 9d and 9g was high for NCI/ADR-RES (89.79 and 39.62 µM, respectively), it is noteworthy that the hybridization also induced a better sensitization of this resistant ovarian adenocarcinoma cell line in comparison to their parental chalcone (>342 and >381 µM, respectively) and DHPM (>332 and >255 µM, respectively).
Hybrid compound 9h was also very potent for breast adenocarcinoma MCF-7 cell line (IC50 value of 5.81 µM) differing from their parental chalcone 4c (IC50 value > 381 µM) and DHPM 8b (IC50 value > 255 µM). Interestingly, different from 9d and 9g, hybrid 9h was not cytotoxic for HaCaT non-tumoral cell line at low concentrations (Table 3). The cytotoxicity for the non-tumoral cell line observed for parental chalcone 4c was reduced with the hybridization process.
Table 3 contains information about the hybrids’ selective index (SI). All of them were selective for breast adenocarcinoma MCF-7, with SI greater than 2, as recommended by the National Cancer Institute [50], suggesting a selective cytotoxic effect for these tumor cells in detriment of healthy ones with hybrid compound 9h presenting a more pronounced SI value (17.33) for MCF-7. This result is very interesting, considering that although substantial progress has been made in cancer chemotherapy, intolerable side effects frequently occur due to the unsatisfactory selectivity of chemotherapeutic agents between tumoral and non-tumoral cells. Hence, the development of new agents with selective chemotherapeutic effects is highly desired [51]. In this regard, hybrid 9h deserves further investigation due to its preference for this tumoral cell line.
The achievement of hybrid compounds from chalcones and dihydropyrimidinones with greater cytotoxicity than their parental compounds is strong evidence of the benefits brought by the molecular hybridization strategy. The fact that hybridization does not always produce greater cytotoxicity than their compounds appears to be more a function of the cell susceptibility of each cell line than the hybrid itself.
The great selectivity of hybrids for MCF-7 breast cancer cells has aroused interest in investigating the action of these molecules on α- and β-estrogen receptors (ERα and ERβ, respectively), highly present in this cell line [52,53], classified as luminal [54], which would exert a blocking effect on the growth of breast cancer tumor. The activation of ERα is responsible for the increase of breast tumor cell proliferation, while ERβ has antiproliferative properties. Therefore, compounds that could act as ERα antagonists and/or ERβ agonists for estrogen-based neoplasia control would have positive effects [55].
To gain preliminary information on this assumption, the flexible molecular docking calculation was performed for the most potent active compounds 9d, 9g, and 9h at the active site of ERα and Erβ. The results show that the compounds 9d, 9g, and 9h were positioned at the same site of an endogenous ligand for ERα (estradiol), with stable interaction, high binding energy and short distance to the amino acids from the pocket (Table 4), which would disturb the action of proliferative molecules, impairing the cell growth, as shown by the cell culture experiments. Besides Thr347, Asp351, and Glu353, binding residues of agonist, the hybrids could bind to Leu 346, indicating more stability to the protein. Moreover, the hybrids docked in the same position as the H3B-9224 antagonist, used as a template for the receptor antagonism, confirming our hypothesis of the participation of ERα receptors as a mechanism of action of hybrids (Figure 3).
On the other hand, the hybrids could bind to the ERβ receptor but not at the active site (Figure 4), not characterizing an agonist or antagonist action, although with high binding energy (Table 5). This data reinforces the antiproliferative activity of the hybrids as an antagonist of ERα receptors only, promoting a reduction of cell viability, and not an agonist of the antiproliferative receptor ERβ.

3. Materials and Methods

3.1. Chemistry

The reagents were purchased from commercial sources and used without further purification, except ethyl acetate and hexanes, which were purified by simple distillation. Column chromatography was performed using a Silica Gel 60 Å (ACROS Organics, 0.035–0.070 mm). The reactions were monitored using thin-layer chromatography (TLC) performed with plates containing silica gel (Merck 60GF245), and the spots were visualized under UV light.
The NMR spectra were recorded using Varian VNMRS 300 spectrometer (1H at 300 MHz and 13C at 75 MHz) or Bruker (1H at 400 MHz and 13C at 100 MHz in CDCl3 or DMSO-d6 as the solvent). The chemical shifts (δ) are reported in ppm units downfield from tetramethylsilane (TMS), which was used as an internal standard. The coupling constants (J) are reported in Hz and refer to peak multiplicities. Attenuated Total Reflection Fourier Transform Infrared (ATR-FTIR) spectra were collected using Bruker Alpha-P equipment. The melting points were obtained on a Buchi Melting Point M-565 apparatus with a non-calibrated thermometer.
The 1H NMR and 13C NMR data and spectra of compounds can be visualized in Supplementary Materials (Figures S1–S41).

3.1.1. General Procedure for Synthesis of Propargyloxy Benzaldehydes (2ae) [43]

In a 50 mL round bottom flask charged with 25 mL of acetone were added 10 mmol of hydroxybenzaldehyde (1ae), 20 mmol (2.764 g) of K2CO3, and 15 mmol (1.60 mL) of propargyl bromide (80% solution in toluene). The resulting solution was stirred under reflux for 2–4 h. After the consumption of aldehyde (monitored by TLC), the crude mixture was filtered, and the volatiles was retired under a vacuum. The products were sufficiently pure (by 1H NMR) to proceed without further purification.

2-(Prop-2-ynil-1-oxy)benzaldehyde (2a)

White solid (m.p. 65 °C, lit. 66–68 °C [43]), 95% yield. 1H NMR (400 MHz, CDCl3) δ 10.49 (s, 1H), 7.87 (dd, 1H, J = 7.8 and 2.0 Hz), 7.57 (ddd, 1H, J = 8.5, 7.3 and 2.0 Hz), 7.15–7.06 (m, 2H), 4.84 (d, 2H, J = 2.3 Hz), 2.57 (t, 1H, J = 2.3 Hz). 13C NMR (100 Hz, CDCl3) δ 189.4, 159.6, 135.6, 128.4, 125.4, 121.6, 113.1, 77.6, 76.4, 56.3.

3-(Prop-2-ynil-1-oxy)benzaldehyde (2b)

Colorless oil [43], 82% yield. 1H NMR (400 MHz, CDCl3) δ 9.98 (s, 1H), 7.53–7.45 (m, 3H), 7.27–7.23 (m, 1H), 4.76 (d, 2H, J = 2.3 Hz), 2.56 (t, 1H, J = 2.3 Hz). 13C NMR (100 Hz, CDCl3) δ 191.7, 157.9, 137.6, 130.0, 123.9, 121.9, 113.4, 77.8, 76.1, 55.8.

4-(Prop-2-ynil-1-oxy)benzaldehyde (2c)

White solid (m.p. 70 °C, lit. 73 °C [43]), 90% yield. 1H NMR (400 MHz, CDCl3) δ 9.89 (s, 1H), 7.87 (d, 2H, J = 8.8 Hz), 7.08 (d, 2H, J = 8.8 Hz), 4.77 (d, 2H, J = 2.5 Hz), 2.58 (t, 1H, J = 2.5 Hz). 13C NMR (100 Hz, CDCl3) δ 190.7, 162.4, 131.9, 130.6, 115.2, 77.5, 76.3, 55.9.

3-Methoxy-4-(prop-2-ynil-1-oxy)benzaldehyde (2d)

White solid (m.p. 90 °C, lit. 86 °C [56]), 90% yield. 1H NMR (400 MHz, CDCl3) δ 9.87 (s, 1H), 7.45 (dd, 1H, J = 8.0 and 1.8 Hz), 7.44 (d, 1H, J = 1.8 Hz), 7.15 (d, 1H, J = 8.0 Hz), 4.86 (d, 2H, J = 2.3 Hz), 3.94 (s, 3H), 2.57 (t, 1H, J = 2.3 Hz). 13C NMR (100 Hz, CDCl3) δ 190.9, 152.1, 150.0, 130.9, 126.2, 112.6, 109.5, 77.5, 76.7, 56.6, 56.0.

4-Methoxy-3-(prop-2-ynil-1-oxy)benzaldehyde (2e)

White solid (m.p. 74 °C, lit. 74–75 °C [57]), 89% yield. 1H NMR (400 MHz, CDCl3) δ 9.85 (s, 1H), 7.53 (d, 1H, J = 1.5 Hz), 7.51 (dd, 1H, J = 8.2 and 1.8 Hz), 7.00 (d, 1H, J = 8.1 Hz), 4.81 (d, 2H, J = 2.3 Hz), 3.95 (s, 3H), 2.54 (t, 1H, J = 2.3 Hz). 13C NMR (75 MHz, CDCl3) δ 190.6, 154.9, 147.3, 129.9, 127.3, 112.0, 110.9, 77.7, 76.4, 56.6, 56.2.

3.1.2. General Procedure for Synthesis of Chalcones (4ae) [44]

In a 25 mL round bottom flask charged with 4 mL of ethanol, was added 2.5 mmol of the propargyloxy benzaldehydes (2ae), and 2.6 mmol of acetophenone (3, 0.3124 g). The mixture was stirred until a clear solution. After, 0.3 mL of a 1M solution of NaOH was added. The reaction was stirred for about 1–2 h and quenched with 10 g of ice water. After, the crude mixture was extracted with 10 mL of CH2Cl2 three times. The organic phases were combined, washed with 20 mL of brine, and dried with MgSO4.
After filtration, the solvent was evaporated under a vacuum. The resultant solid product was purified by silica gel column chromatography (hexane/ethyl acetate 80/20, v/v).

1-Phenyl-3-(2-prop-2-ynil-1-oxyphenyl)-(2E)-propen-1-one (4a)

Yellow solid, (m.p. 64 °C, lit. 63–64.6 °C [58]), 82% yield. 1H NMR (400 MHz, CDCl3) δ 8.10 (d, 1H, J = 15.9 Hz), 8.06-8.01 (m, 2H), 7.67 (d, 1H, J = 15.9 Hz), 7.65 (dd, 1H, J = 7.8 and 1.8 Hz), 7.61–7.54 (m, 1H), 7.52–7.46 (m, 2H), 7.38 (ddd, 1H, J = 8.5, 7.2, and 1.8 Hz), 7.07–7.03 (m, 2H), 4.80 (d, 2H, J = 2.5 Hz), 2.56 (t, 1H, J = 2.5 Hz). 13C NMR (100 MHz, CDCl3) δ 191.0, 158.7, 140.0, 138.4, 132.6, 131.5, 129.5, 128.5, 124.5, 123.3, 121.7, 112.7, 78.2, 76.0, 56.2.

1-Phenyl-3-(3-prop-2-ynil-1-oxyphenyl)-(2E)-propen-1-one (4b)

Yellow solid (m.p. 54 °C, Iit. 55 °C [59]), 67% yield. 1H NMR (400 MHz, CDCl3) δ 8.04–7.99 (m, 2H), 7.77 (d, 1H, J = 15.6 Hz), 7.62–7.57 (m, 1H), 7.51 (d, 2H, J = 15.6 Hz), 7.53–7.48 (m, 2H), 7.39–7.33 (m, 1H), 7.31–7.27 (m, 1H), 7.25–7.23 (m, 1H), 7.04 (ddd, 1H, J = 8.1, 2.5 and 0.8 Hz), 4.74 (d, 2H, J = 2.3 Hz), 2.56 (t, 1H, J = 2.3 Hz). 13C NMR (100 MHz, CDCl3) δ 190.5, 144.5, 138.1, 136.3, 132.9, 130.0, 128.6, 128.5, 122.5, 121.9, 117.1, 114.6, 78.2, 75.9, 55.9.

1-Phenyl-3-(4-prop-2-ynil-1-oxyphenyl)-(2E)-propen-1-one (4c)

Yellow solid (m.p. 72 °C, 70–72 °C [60]), 71% yield. 1H NMR (400 MHz, CDCl3) δ 8.04–7.98 (m, 2H), 7.78 (d, 1H, J = 15.6 Hz), 7.61 (d, 2H, J = 8.6 Hz), 7.61–7.54 (m, 1H), 7.50 (t, 2H, J = 7.0 Hz), 7.43 (d, 1H, J = 15.6 Hz), 7.02 (d, 2H, J = 8.6 Hz), 4.74 (d, 2H, J = 2.3 Hz), 2.55 (t, 1H, J = 2.3 Hz). 13C NMR (100 MHz, CDCl3) δ 190.6, 159.5, 144.4, 138.4, 132.6, 130.2, 128.6, 128.5, 128.4, 120.3, 115.3, 78.0, 76.0, 55.9.

1-Phenyl-3-(3-methoxy-4-prop-2-ynil-1-oxyphenyl)-(2E)-propen-1-one (4d)

Yellow solid (m.p. 110 °C, lit. 108 °C [34]), 60% yield. 1H NMR (400 MHz, CDCl3) δ 8.01 (d, 2H, J = 7.3 Hz), 7.76 (d, 1H, J = 15,6 Hz), 7.58 (t, 1H, J = 7.3 Hz), 7.50 (t, 2H, J = 7.3 Hz), 7.41 (d, 1H, J = 15.8 Hz), 7.23 (dd, 1H, J = 8.3 and 1,5 Hz), 7.18 (d, 1H, J = 1.5 Hz), 7.06 (d, 1H, J = 8.3 Hz), 4.82 (d, 2H, J = 2.3 Hz), 3.94 (s, 3H), 2.55 (t, 1H, J = 2.3 Hz). 13C NMR (75 MHz, CDCl3) δ 190.5, 149.7, 149.0, 144.7, 138.3, 132.6, 129.0, 128.5, 128.4, 122.5, 120.5, 113.7, 110.7, 77.9, 76.2, 56.6, 55.94.

1-Phenyl-3-(4-methoxy-3-prop-2-ynil-1-oxyphenyl)-(2E)-propen-1-one (4e)

Yellow solid (m.p. 103 °C), 62% yield. 1H NMR (40 MHz, CDCl3) δ 8.03–8.00 (m, 2H), 7.77 (d, 1H, J = 15.6 Hz), 7.62–7.56 (m, 1H), 7.54–7.48 (m, 2H), 7.40 (d, 1H, J = 15.6 Hz), 7.35 (d, 1H, J = 2.0 Hz), 7.32–7.28 (m, 1H), 6.93 (d, 1H, J = 8.3 Hz), 4.83 (d, 2H, J = 2.3 Hz), 3.93 (s, 3H), 2.57 (t, 1H, J = 2.3 Hz). 13C NMR (100 Hz, CDCl3) δ 190.5, 151.9, 146.8, 144.7, 138.3, 132.6, 128.5, 128.4, 127.7, 124.1, 120.2, 113.5, 111.6, 78.1, 76.3, 56.8, 55.9. ATR-FTIR (νmax, cm−1) 3277, 3001, 2907, 2831, 2130, 1658, 1513, 1013. HRMS (ESI) m/z calc. for [C19H16O3 + Na]+: 315.0992; obs. 315.0993.

3.1.3. General Procedure for the Synthesis of Chloro-DHPMs (7a,b) [45]

In a 50 mL round-bottom flask, were added, in this order, 5 mmol of the benzaldehydes 1f,g, 5 mmol (0.503 mL) of ethyl 4-chloroacetoacetate (6), 7.5 mmol (0.4505 g) of urea (7) and 12.5 mL of acetic acid. The mixture was stirred under reflux for 48 h. After cooling, the solution was slowly poured into an Erlenmeyer flask containing 100 mL of water, and a solid appeared. The solid was filtrated on a Büchner funnel and washed with water (25 mL), saturated NaHCO3 solution (25 mL), and water again (25 mL). The solid was dried under reduced pressure.

Ethyl 6-(2-chloromethyl)-4-phenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (7a)

Pale yellow solid (m.p. 200 °C, lit. 226 °C [50]), 72% yield. 1H NMR (300 MHz, DMSO-d6) δ 9.24 (ls, 1H), 7.63 (ls, 1H), 7.35–7.21 (m, 5H), 5.22 (d, 1H, J = 2.9 Hz), 4.77 (d, 1H, J = 10,6 Hz), 4.61 (d, 1H, J = 10,6 Hz), 4.06 (q, 2H, J = 7.0 Hz), 1.12 (t, 1H, J = 7.0 Hz). 13C NMR (100 MHz, DMSO-d6) δ 164.2, 152.1, 146.0, 144.0, 128.5, 127.6, 126.3, 101.8, 60.0, 53.9, 39.2, 13.9.

Ethyl 6-chloromethyl-4-(3,4,5-trimethoxyphenyl)-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (7b)

Dark yellow solid (m.p. 244 °C), 70% yield. 1H NMR (400 MHz, DMSO-d6) δ 9.50 (ls, 1H), 7.82 (ls, 1H), 6.57 (s, 2H), 5.15 (d, 1H, J = 3.3 Hz), 4.80 (d, 1H, J = 10.6 Hz), 4.64 (d, 1H, J = 10.6 Hz), 4.08 (q, 2H, J = 7.0 Hz), 3.73 (s, 6H), 3.64 (s, 3H), 1.14 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, DMSO-d6) δ 172.1, 164.3, 152.9, 152.1, 146.3, 139.4, 137.0, 103.5, 101.4, 60.0, 55.8, 53.8, 39.4, 21.1, 14.0. ATR-FTIR (νmax, cm−1 3318, 3210, 3098, 2936, 1686, 1643, 1220, 1123, 737. HRMS (ESI) m/z calc. for [C17H21ClN2O6 + Na]+: 407.0980; obs. 407.0986.

3.1.4. General Procedure for the Synthesis of Azido-DHPMs (8a,b) [46]

A 50 mL round bottom flask was charged with 7.5 mL acetone and 3 mL deionized water. Next, 3 mmol of chloro-DHPM (7a,b) and 3.6 mmol (0.2340 g) of sodium azide were added at room temperature. The mixture was heated at 60 °C until the consumption of the Cl-DHPM (monitored by TLC). Acetone was evaporated under a vacuum, and the resulting solid was washed with water (25 mL), filtered on a Büchner funnel, and dried at reduced pressure.

Ethyl 6-(azidomethyl)-4-phenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (8a)

Yellow solid (m.p. 111 °C, lit. 110 [61]), 88% yield. 1H NMR (400 MHz, DMSO-d6) δ 9.47 (ls, 1H), 7.88 (ls, 1H), 7.37–7.31 (m, 2H), 7.29–7.23 (m, 3H), 5.21 (d, 1H, J = 3.3 Hz), 4.86 (d, 1H, J = 12.8 Hz), 4.35 (d, 1H, J = 12.8 Hz), 4.03 (q, 2H, J = 7.0 Hz), 1.11 (t, 1H, J = 7.0 Hz). 13C NMR (100 MHz, DMSO-d6) δ 164.7, 152.0, 144.4, 144.1, 128.6, 127.7, 126.4, 102.4, 60.1, 54.1, 48.0, 13.9.

Ethyl 6-(azidomethyl)-4-(3,4,5-trimethoxyphenyl)-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (8b)

Pale yellow solid (mp 177 °C), 91% yield. 1H NMR (400 MHz, DMSO-d6) δ 9.45 (ls, 1H), 7.83 (d, 1H, J = 2.8 Hz), 6.57 (s, 2H), 5.19 (d, 1H, J = 3.0 Hz), 4.50 (d, 1H, J = 12.8 Hz), 4.31 (d, 1H, J = 12.8 Hz), 4.07 (qd, 2H, J = 7.0 and 1.8 Hz), 3.73 (s, 6H), 3.63 (s, 3H), 1.13 (t, 1H, J = 7.0 Hz). 13C NMR (100 MHz, DMSO-d6) δ 172.1, 164.3, 152.9, 152.1, 146.3, 139.4, 137.0, 103.5, 101.4, 60.0, 55.8, 53.8, 39.4, 21.1, 14.0. ATR-FTIR (νmax, cm−1) 3321, 3209, 3116, 2939, 2095, 1683, 1224, 1125. HRMS (ESI) m/z calc. for [C17H21N5O6 + Na]+: 414.1384, obs. 414.1386.

3.1.5. General Procedure for the Synthesis of Chalcone-DHPM Hybrids (9aj) [47]

In a 25 mL round bottom flask was placed 2 mL of CH2Cl2 and 2 mL of water. Then, 0.2 mmol of azido-DHPM (8a,b) was added, followed by 0.2 mmol of propargyloxy chalcone (4ae), 0.02 mmol (0.0050 g) of CuSO4·5H2O and 0.02 mmol (0.0040 g) of sodium ascorbate. The reaction was stirred, at room temperature, until consumption of starting materials (monitored by TLC). After, 4 mL of EDTA aqueous solution (0.1 M) was added and steered for 5 min. The crude aqueous mixture was extracted with 5 mL of CH2Cl2 three times. The organic phases were combined, washed with 10 mL of saturated NaCl solution, and dried over MgSO4. After filtration, the solvent was removed under reduced pressure. The resulting solid was purified by silica gel column chromatography (ethyl acetate/hexanes, 70/30, v/v).

Ethyl 6-((4-((2-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-phenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9a)

Yellow solid (m.p. 109 °C), 75% yield. 1H NMR (400 MHz, CDCl3) δ 9.00 (ls, 1H), 8.09 (d, 1H, J = 15,6 Hz), 8.04 (s, 1H), 7.96 (d, 2H, J = 7.3 Hz), 7.63 (d, 1H, J = 7.5 Hz), 7.62 (d, 1H, J = 15.8 Hz), 7.57 (t, 1H, J = 7.3 Hz), 7.47 (d, 2H, J = 7.3 Hz), 7.37 (d, 1H, J = 7.5 Hz), 7.21 (s, 5H), 7.08 (d, 1H, J = 8.3 Hz), 7.02 (t, 1H, J = 7.5 Hz), 6.33 (ls, 1H), 5.85 (d, 1H, J = 14.6 Hz), 5.59 (d, 1H, J = 14.6 Hz), 5.38 (d, 1H, J = 2.5 Hz), 5.29 (s, 2H), 4.10 (q, 2H, J = 7.0 Hz), 1.13 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.8, 164.7, 157.3, 152.9, 143.6, 142.8, 141.2, 140.0, 138.1, 132.8, 131.9, 129.0, 128.7, 128.5, 128.5, 128.1, 126.6, 124.2, 124.1, 122.4, 121.3, 112.6, 104.5, 62.6, 60.8, 55.6, 47.7, 13.9. ATR-FTIR (νmax, cm−1) 3238, 3143, 2935, 1688, 1653, 1217, 1098, 1005. HRMS (ESI) m/z calc. for [C32H29N5O5 + Na]+: 586.2061; obs. 586.2061.

Ethyl 6-((4-((3-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3- triazol-1-yl)methyl)-4-phenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9b)

Yellow solid (m.p. 91 °C), 70% yield. 1H NMR (400 MHz, CDCl3) δ 8.35 (ls, 1H), 8.02 (d, 1H, J = 7.3 Hz), 7.91 (s, 1H), 7.76 (d, 1H, J = 15.6 Hz), 7.59 (t, 1H, J = 7.3 Hz), 7.55–7.47 (m, 3H), 7.36–7.21 (m, 7H), 7.06–7.01 (m, 1H), 6.08 (ls, 1H), 5.82 (d, 1H, J = 14.6 Hz), 5.64 (d, 1H, J = 14.9 Hz), 5.40 (d, 1H, J = 2.5 Hz), 5.22 (s, 2H), 4.10 (q, 2H, J = 7.0 Hz), 1.15 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.4, 164.8, 158.5, 152.4, 144.5, 142.7, 141.4, 138.0, 136.3, 132.8, 130.0, 128.8, 128.6. 128.5, 128.3, 126.6, 124.0, 122.4, 121.7, 116.9, 114.4, 104.4, 61.9, 60.9, 55.6, 47.9, 13.9. ATR-FTIR (νmax, cm−1) 3226, 3114, 2956, 1693, 1656, 1221, 1094, 1014. HRMS (ESI) m/z calc. for [C32H29N5O5 + Na]+: 586.2061; obs. 586.2064.

Ethyl 6-((4-((4-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-phenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9c)

Yellow solid (m.p. 111 °C), 78% yield. 1H NMR (400 MHz, CDCl3) δ 8.43 (ls, 1H), 8.03-7.98 (m, 2H), 7.90 (s, 1H), 7.77 (d, 1H, J = 15.6 Hz), 7.61–7.55 (m, 3H), 7.52–7.47 (m, 2H), 7.42 (d, 1H, J = 15.6 Hz), 7.29–7.20 (m, 5H), 7.01 (d, 1H, J = 8.8 Hz), 6.10 (ls, 1H), 5.77 (d, 1H, J = 14.6 Hz), 5.65 (d, 1H, J = 14.6 Hz), 5.40 (d, 1H, J = 2.8 Hz), 5.22 (s, 2H), 4.09 (q, 2H, J = 7.0 Hz), 1.15 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.5, 164.8, 160.2, 152.4, 144.4, 143.7, 142.7, 141.4, 138.4, 132.6, 130.2, 128.8, 128.7. 128.4, 128.3, 128.1, 126.6, 124.1, 120.0, 115.2, 104.4, 61.9, 60.9, 55.6, 49.0, 13.9. ATR-FTIR (νmax, cm−1) 3238, 3131, 2957, 1695, 1643, 1211, 1171, 1016. HRMS (ESI) m/z calc. For [C32H29N5O5 + Na]+: 586.2061; obs.586.2062.

Ethyl 6-((2-methoxy-4-((4-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-phenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9d)

Yellow solid (m.p. 107 °C), 90% yield. 1H NMR (400 MHz, CDCl3) δ 8.48 (ls, 1H), 8.03–7.99 (m, 2H), 7.95 (s, 1H), 7.74 (d, 1H, J = 15.6 Hz), 7.61–7.56 (m, 1H), 7.58 (t, 2H, J = 7.3 Hz), 7.39 (d, 1H, J = 15.6 Hz), 7.29–7.18 (m, 6H), 7.14 (d, 1H, J = 1.7 Hz), 7.07 (d, 1H, J = 8.3 Hz), 6.06 (ls, 1H), 5.82 (d, 1H, J = 14.6 Hz), 5.60 (d, 1H, J = 14.6 Hz), 5.38 (d, 1H, J = 2.5 Hz), 5.28 (s, 2H), 4.09 (q, 2H, J = 7.0 Hz), 3.87 (s, 3H), 1.14 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.6, 164.8, 152.4, 149.9, 149.5, 144.9, 143.7, 142.8, 141.4, 138.4, 132.7, 128.8, 128.6. 128.5, 128.3, 126.7, 124.5, 122.8, 120.4, 113.6, 110.7, 104.3, 62.8, 60.9, 55.9, 55.7, 47.9, 14.0. ATR-FTIR (νmax, cm−1) 3299, 3133, 2937, 1694, 1651, 1221, 1096, 1013. HRMS (ESI) m/z calc. for [C33H31N5O6 + Na]+: 616.2167; obs. 616.2175.

Ethyl 6-((3-methoxy-4-((3-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-phenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9e)

Yellow solid (m.p. 112 °C), 75% yield. 1H NMR (300 MHz, CDCl3) δ 8.39 (ls, 1H), 8.04 (d, 2H, J = 7.2 Hz), 7.96 (s, 1H), 7.76 (d, 1H, J = 15.8 Hz), 7.61–7.54 (m, 1H), 7.54–7.47 (m, 2H), 7.47–7.40 (m, 2H), 7.31–7.20 (m, 6H), 6.89 (d, 1H, J = 8.5 Hz), 6.03 (ls, 1H), 5.85 (d, 1H, J = 14.5 Hz), 5.59 (d, 1H, J = 14.5 Hz), 5.38 (d, 1H, J = 2.6 Hz), 5.30 (s, 2H), 4.09 (q, 2H, J = 7.0 Hz), 3.85 (s, 3H), 1.14 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.4, 164.7, 152.5, 15.8, 147.8, 144.6, 143.6, 142.8, 141.4, 138.3, 132.5, 128.7, 128.5, 128.4, 128.1, 127.8. 126.6, 124.5, 124.1, 120.1, 113.1, 111.5, 104.3, 62.9, 60.7, 55.8, 55.5, 47.8, 13.8. ATR-FTIR (νmax, cm−1) 3232, 3114, 2937, 1693, 1650, 1226, 1094, 1012. HRMS (ESI) m/z calc. for [C33H31N5O6 + Na]+: 616.2167; obs. 616.2174.

6-((4-((2-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-(3,4,5-trimethoxyphenyl)-3,4-dihydropyrimidin-(1H)-2-one-5 carboxylate (9f)

Yellow solid (m.p. 107 °C), 72% yield. 1H NMR (300 MHz, CDCl3) δ 9.39 (ls, 1H), 8.14 (d, 1H, J = 15,6 Hz), 8.14 (s, 1H), 7.96 (d, 2H, J = 7.5 Hz), 7.65 (d, 1H, J = 7.0 Hz), 7.58 (d, 1H, J = 15.6 Hz), 7.56 (t, 1H, J = 7.8 Hz), 7.47 (t, 2H, J = 7.8 Hz), 7.39 (t, 1H, J = 7.3 Hz), 7.07 (d, 1H, J = 8.3 Hz), 7.03 (t, 1H, J = 7.5 Hz), 6.56 (ls, 1H), 6.41 (s, 2H), 5.91 (d, 1H, J = 14.1 Hz), 5.57 (d, 1H, J = 14.3 Hz), 5.34 (d, 1H, J = 2.3 Hz), 5.27 (s, 2H), 4.11 (q, 2H, J = 7.0 Hz), 3.77 (s, 3H), 3.66 (s, 6H), 1.15 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.7, 164.7, 157.2, 153.3, 153.0, 143.6, 141.3, 139.8, 138.4, 138.0, 137.6, 132.9, 132.1, 128.6, 128.6, 128.5, 124.4, 124.0, 122.1, 121.5, 112.6, 104.2, 103.5, 62.8, 60.8, 60.7, 55.9, 55.9, 47.8, 14.0. ATR-FTIR (νmax, cm–1) 3214, 3097, 2936, 1693, 1649, 1217, 1126, 1000. HRMS (ESI) m/z calc. for [C35H35N5O8 + Na]+: 676.2378; obs. 676.2380.

Ethyl 6-((4-((3-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-(3,4,5-trimethoxyphenyl)-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9g)

Yellow solid (m.p. 103 °C), 70% yield. 1H NMR (400 MHz, CDCl3) δ 8.42 (ls, 1H), 8.05–7.99 (m, 2H), 7.97 (s, 1H), 7.76 (d, 1H, J = 15.6 Hz), 7.62–7.56 (m, 3H), 7.56–7.48 (m, 3H), 7.36–7.30 (t, 1H, J = 8.1 Hz), 7.27–7.23 (m, 2H), 7.05–7.01 (m, 1H), 6.43 (s, 1H), 6.07 (ls, 1H), 5.91 (d, 1H, J = 14.4 Hz), 5.58 (d, 1H, J = 14.4 Hz), 5.37 (d, 1H, J = 2.5 Hz), 5.21 (s, 2H), 4.13 (q, 2H, J = 7.1 Hz), 3.79 (s, 3H), 3.74 (6H, s), 1.18 (t, 3H, J = 7.1 Hz). 13C NMR (100 MHz, CDCl3) δ 190.4, 164.7, 158.5, 153.4, 152.4, 144.4, 143.9, 141.5, 138.3, 138.0, 137.8, 136.4, 132.8, 130.0, 128.6, 128.5, 124.1, 122.5, 121.7, 116.9, 114.3, 104.1, 103.4, 61.9, 60.9, 60.7, 56.0, 55.8, 48.1, 14.0. ATR-FTIR (νmax, cm−1) 3228, 3102, 2939, 1693, 1653, 1222, 1093, 1013. HRMS (ESI) m/z calc. for [C35H35N5O8 + Na]+: 672.2378; obs. 672.2373.

Ethyl 6-((4-((4-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-(3,4,5-trimethoxyphenyl)-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9h)

Yellow solid (m.p. 109 °C), 62% yield. 1H NMR (400 MHz, CDCl3) δ 8.56 (ls, 1H), 8.05–8.01 (m, 2H), 7.98 (s, 1H), 7.79 (d, 1H, J = 15.6 Hz), 7.63–7.58 (m, 1H), 7.60 (t, 2H, J = 8.8 Hz), 7.52 (t, 2H, J = 7.0 Hz), 7.44 (d, 1H, J = 15.6 Hz), 7.02 (d, 2H, J = 8.8 Hz), 6.44 (s, 1H), 6.17 (ls, 1H), 5.94 (d, 1H, J = 14.3 Hz), 5.57 (d, 1H, J = 14.3 Hz), 5.38 (d, 1H, J = 2.8 Hz), 5.22 (s, 2H), 4.14 (q, 2H, J = 7.0 Hz), 3.82 (s, 3H), 3.76 (s, 6H) 1.19 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.5, 164.7, 160.1, 153.5, 152.3, 144.4, 143.7, 141.5, 138.3, 138.2, 137.8, 132.6, 130.2, 128.5, 128.4, 128.2, 124.2, 120.1, 115.1, 104.1, 103.4, 61.8, 60.9, 60.8, 56.0, 55.8, 48.1, 14.0. ATR-FTIR (νmax, cm−1) 3229, 3104, 2936, 2831, 1696, 1653, 1217, 1119, 1000. HRMS (ESI) m/z calc. for [C 35 H 35 N 5 O 8 + Na]+: 676.2378, obs. 676.2377.

Ethyl 6-((3-methoxy-4-((4-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy)methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-(3,4,5-trimethoxyphenyl-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9i)

Yellow solid (m.p. 102 °C), 90% yield. 1H NMR (400 MHz, CDCl3) δ 8.49 (ls, 1H), 8.04–8.00 (m, 2H), 7.97 (s, 1H), 7.74 (d, 1H, J = 15.6 Hz), 7.61–7.56 (m, 1H), 7.54–7.48 (m, 2H), 7.40 (d, 1H, J = 15.6 Hz), 7.20 (dd, 1H, J = 8.4 and 1.9 Hz), 7.14 (d, 1H, J = 1.8 Hz), 7.06 (d, 1H, J = 8.3 Hz), 6.42 (s, 2H), 6.08 (ls, 1H), 5.85 (d, 1H, J = 14.5 Hz), 5.59 (d, 1H, J = 14.5 Hz), 5.34 (d, 1H, J = 2.5 Hz), 5.26 (s, 2H), 4.12 (q, 2H, J = 7.0 Hz), 3.88 (s, 3H), 3.80 (s, 3H), 3.74 (s, 6H), 1.17 (t, 3H, J = 7.0 Hz). 13C NMR (100 MHz, CDCl3) δ 190.5, 164.7, 153.4, 152.4, 149.8, 149.5, 144.7, 143.5, 141.5, 138.3, 138.3, 137.8, 132.6, 128.7, 128.5, 128.4, 124.7, 128.5, 128.4, 124.7, 122.8, 120.4, 113.6, 110.7, 104.1, 103.4, 62.7, 60.8, 60.7, 56.0, 55.8, 48.0. ATR-FTIR (νmax, cm−1) 3246, 3107, 2929, 2851, 1696, 1653, 1222, 1125, 1000. HRMS (ESI) m/z calc. for [C36H37N5O9 + Na]+: 706.2483; obs. 706.2482.

Ethyl 6-((3-methoxy-4-((3-(3-oxo-3-phenylprop-1-(E)-en-1-yl)phenoxy) methyl)-(1H)-1,2,3-triazol-1-yl)methyl)-4-(3,4,5-trimethoxyphenyl)-3,4-dihydropyrimidin-(1H)-2-one-5-carboxylate (9j)

Yellow solid (m.p. 114 °C), 85% yield. 1H NMR (400 MHz, CDCl3) δ 8.43 (ls, 1H), 8.03–7.98 (m, 2H), 7.90 (s, 1H), 7.77 (d, 1H, J = 15.6 Hz), 7.61–7.55 (m, 1H), 7.53–7.47 (m, 2H), 7.45–7.38 (m, 2H), 7.25 (dd, 1H, J = 8.3 and 2.0 Hz), 6.89 (d, 1H, J = 8.3 Hz), 6.43 (s, 2H), 6.10 (ls, 1H), 5.84 (d, 1H, J = 14.4 Hz), 5.61 (d, 1H, J = 14.4 Hz), 5.34 (d, 1H, J = 2.5 Hz), 5.27 (s, 2H), 4.12 (q, 2H, J = 7.2 Hz), 3.85 (s, 3H), 3.79 (s, 3H), 3.73 (s, 6H), 1.17 (t, 3H, J = 7.2 Hz). 13C NMR (100 MHz, CDCl3) δ 190.5, 164.8, 153.4, 152.3, 151.9, 147.8, 144.7, 143.8, 141.6, 138.4, 138.3, 137.7, 132.7, 128.6, 128.5, 127.9, 124.7, 124.1, 120.2, 113.2, 111.6, 104.1, 103.4, 63.0, 60.9, 60.8, 56.1, 55.9, 55.9, 48.0, 14.1. ATR-FTIR (νmax, cm−1) 3235, 3116, 2940, 1690, 1649, 1221, 1122, 1013. HRMS (ESI) m/z calc. for [C36H37N5O9 + Na]+: 706.2483; obs. 706.2473.

3.2. Cell Culture and MTT Assay

The cytotoxic activity of compounds was evaluated in a panel of commercial human cell lines purchased from the American Type Culture Collection–ATCC: five tumoral (U-251–brain, glioblastoma; MCF-7–breast, adenocarcinoma; NCI/ADR-RES–ovary, multidrug-resistant adenocarcinoma; OVCAR-3–ovary, adenocarcinoma, and HT-29–colorectal, adenocarcinoma) and one non-tumoral (HaCaT–keratinocyte).
For this assay, the colorimetric method 3-(4,5-dimethylthiazol-2-yl)2,5-diphenyltetrazolium bromide, MTT (Sigma-Aldrich, Burlington, MA, USA) was used to indirectly evaluate cell viability by the mitochondrial enzymatic activity of living cells [62]. The cell suspensions were prepared in RPMI-1640 (Sigma-Aldrich, Burlington, MA, USA) medium containing 5% FBS (LGC Biotecnologia, Cotia, SP, Brazil) and 1% PS (LGC Biotecnologia, Cotia, SP, Brazil). One hundred μL of cell suspension containing 5000 cells were inoculated per well into 96-well plates and incubated for 24 h at 37 °C in a 5% CO2 atmosphere and humidity. After 24 h, samples were diluted in DMSO and added to the cells at concentrations ranging from 2.5–381 μM (100 μL/well) in triplicate and then incubated for 48 h at 37 °C in a 5% CO2 atmosphere and humidity. As a positive control, the chemotherapy drug doxorubicin hydrochloride, doxo (Eurofarma, São Paulo, SP, Brazil), was used at concentrations of 0.3–18.4 μM (100 μL/well) in triplicate. The final DMSO concentration (less than 0.25%) did not affect cell viability [63]. After 48 h of treatment, the treated cells were then stained with MTT. The absorbance data were analyzed and compiled in the graphs plotting the percentage of viable cells with the sample concentration. The IC50 (half maximal inhibitory concentration) values were calculated, which refers to the concentration of samples required to decrease in 50% cell viability. Hybrids and their parental molecules were also evaluated for the selectivity index (SI), which reflects the differential cytotoxicity of a compound against tumor and normal cells. The greater the SI value of a compound, the more selective it is. This study obtained SI from the following formula: IC50 of the non-tumoral cell line (HaCaT)/IC50 of tumor cell lines. For this analysis, a SI value greater than or equal to 2.0 was adopted as significant, as previously described by the NCI, USA (National Cancer Institute, Bethesda, MD, USA) [50].

3.3. Molecular Docking

The hybrids were drawn, converted to a PDB file, and then set to rotatable bonds to get flexible docking. ER receptors α and β were selected in the Protein Data Bank (PDB) by their resolution (1.68 and 2.20 Å, respectively) and binding to an antagonist and agonist. For ER-α, it was chosen PDB code 6CHZ and considered the binding pocket of three residues (Thr347, Asp351, and Glu353), the position of the endogenous agonist. For Erβ, the protein 3OLS PDB code was selected and considered the binding pocket Glu305, Arg346, and Phe356. Moreover, hybrids were superimposed onto co-crystallized proteins containing endogenous agonists to check the positioning.
Missing atoms, chain breaks, and water molecules were removed, and hydrogens were added for a pH of 7.0. The grid was centered within the protein’s active site (subunit A) with a size of 22 × 24 × 28 Å in the x, y, and z axes, respectively.
Ligands and proteins were subjected to analysis by AutoDock Vina to get the binding energy (kcal/mol) and to Chimera 1.15 software for distances calculation, and amino acid binds determination.

4. Conclusions

A series of 10 hybrid molecules based on chalcones and dihydropyrimidinones were designed and readily synthesized using simple reactions such as the Claisen–Schmidt condensation, multicomponent Biginelli reaction, and copper-catalyzed Huisgen reaction. Different hybrids showed low IC50 values for different cancer cell lines, such as ovary adenocarcinoma (including one resistant), glioblastoma, and breast adenocarcinoma cell lines. It is noteworthy that all hybrids exhibited powerful cytotoxicity against luminal breast adenocarcinoma MCF-7.
The cytotoxicity of parentals (chalcone and dihydropyrimidinone, respectively) was always smaller than the cytotoxicity of the most potent hybridized compounds 9d, 9g, and 9h, suggesting that potency may be the result of molecular hybridization. Additionally, the hybrids were more selective to breast cancer cell lines than healthy cell lines, which is desirable for a drug candidate. In this sense, compound 9h showed the highest SI, deserving a further complementary study.
In silico studies denoted that the hybrid compounds 9d, 9g, and 9h are positioned on the active site of the ERα or in its surroundings, acting as antagonist molecules that can, in turn, disrupt or hinder the cell proliferation process. These findings deeply encouraged us to proceed with further investigations to elucidate the mechanism of action against luminal breast cancers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ddc1010002/s1, Figure S1: 1H NMR (400 MHz, CDCl3) of 4a. Figure S2: 13C NMR (100 MHz, CDCl3) of 4a. Figure S3: 1H NMR (400 MHz, CDCl3) of 4b. Figure S4: 13C NMR (100 MHz, CDCl3) of 4b. Figure S5: 1H NMR (400 MHz, CDCl3) of 4c. Figure S6: 13C NMR (400 MHz, CDCl3) of 4c. Figure S7: 1H NMR (400 MHz, CDCl3) of 4d. Figure S8: 13C NMR (75 MHz, CDCl3) of 4d. Figure S9: 1H NMR (400 MHz, CDCl3) of 4e. Figure S10: 13C NMR (400 MHz, CDCl3) of 4e. Figure S11: 1H NMR (400 MHz, DMSO-d6) of 7a. Figure S12: 13C NMR (100 MHz, DMSO-d6) of 7a. Figure S13: 1H NMR (400 MHz, DMSO-d6) of 7b. Figure S14: 13C NMR (100 MHz, DMSO-d6) of 7b. Figure S15: 1H NMR (400 MHz, DMSO-d6) of 7a. Figure S16: 13C NMR (100 MHz, DMSO-d6) of 8a. Figure S17: 1H NMR (100 MHz, DMSO-d6) of 8b. Figure S18: 13C NMR (100 MHz, DMSO-d6) of 8b. Figure S19: 1H NMR (400 MHz, CDCl3) of 9a. Figure S20: 13C NMR (100 MHz, CDCl3) of 9a. Figure S21: 1H NMR (400 MHz, CDCl3) of 9b. Figure S22: 13C NMR (100 MHz, CDCl3) of 9b. Figure S23: 1H NMR (400 MHz, CDCl3) of 9c. Figure S24: 13C NMR (100 MHz, CDCl3) of 9c. Figure S25: 1H NMR (400 MHz, CDCl3) of 9d. Figure S26: 13C NMR (100 MHz, CDCl3) of 9d. Figure S27: 1H NMR (400 MHz, CDCl3) of 9e. Figure S28: 13C NMR (100 MHz, CDCl3) of 9e. Figure S29: 1H NMR (400 MHz, CDCl3) of 9f. Figure S30: 13C NMR (100 MHz, CDCl3) of 9f. Figure S31: 1H NMR (400 MHz, CDCl3) of 9g. Figure S32: 13C NMR (100 MHz, CDCl3) of 9g. Figure S33: 1H NMR (400 MHz, CDCl3) of 9h. Figure S34: 13C NMR (100 MHz, CDCl3) of 9h. Figure S35: 1H NMR (400 MHz, CDCl3) of 9i. Figure S36: 13C NMR (100 MHz, CDCl3) of 9i. Figure S37: 1H NMR (400 MHz, CDCl3) of 9j. Figure S38: 13C NMR (100 MHz, CDCl3) of 9j. Figure S39: Cell viability of parentals 4d and 8a and hybrid 9d. Figure S40: Cell viability of parentals 4b and 8b and hybrid 9g. Figure S41: Cell viability of parentals 4c and 8b and hybrid 9h.

Author Contributions

Conceptualization, D.R.; methodology, E.B.M. and C.A.d.L.; formal analysis and investigation, D.R., M.G.M.D., G.B.L. and J.M.S.; writing—original draft preparation, E.B.M. and D.R.; writing—review and editing, D.R. and G.B.L.; funding acquisition, D.R. and G.B.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the Fundação de Apoio à Pesquisa do Estado do Rio Grande do Sul (FAPERGS, grant no. 19/2551-0001767-7), Fundação de Apoio à Pesquisa do Estado de São Paulo (FAPESP, grant no. 2016/06137-5) and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq grant no. 312452/2017-9). E.B.M. and C.A.d.L. acknowledge the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) for the fellowships.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bérubé, G. An Overview of Molecular Hybrids in Drug Discovery. Expert Opin. Drug Discov. 2016, 11, 281–305. [Google Scholar] [CrossRef] [PubMed]
  2. Talevi, A. Multi-target pharmacology: Possibilities and limitations of the “skeleton key approach” from a medicinal chemist perspective. Front. Pharmacol. 2015, 6, 205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Morphy, R.; Rankovic, Z. Designed Multiple Ligands. An Emerging Drug Discovery Paradigm. J. Med. Chem. 2005, 48, 6523–6543. [Google Scholar] [CrossRef] [PubMed]
  4. Nepali, K.; Sharma, S.; Sharma, M.; Bedi, P.M.S.; Dhar, K.L. Rational Approaches, Design Strategies, Structure Activity Relationship and Mechanistic Insights for Anticancer Hybrids. Eur. J. Med. Chem. 2014, 77, 422–487. [Google Scholar] [CrossRef] [PubMed]
  5. Decker, M. (Ed.) Design of Hybrid Molecules for Drug Development; Elsevier: Cambridge, MA, USA, 2017. [Google Scholar]
  6. Bansal, Y.; Silakari, O. Multifunctional compounds: Smart molecules for multifactorial diseases. Eur. J. Med. Chem. 2014, 76, 31–42. [Google Scholar] [CrossRef]
  7. Tsogoeva, S.B. Recent Progress in the Development of Synthetic Hybrids of Natural or Unnatural Bioactive Compounds for Medicinal Chemistry. Mini Rev. Med. Chem. 2010, 10, 773–793. [Google Scholar] [CrossRef]
  8. Tietze, L.F.; Bell, H.P.; Chandrasekhar, S. Natural Product Hybrids as New Leads for Drug Discovery. Angew. Chem. Int. Ed. 2003, 42, 3996–4028. [Google Scholar] [CrossRef]
  9. Vendrusculo, V.; de Souza, V.P.; Fontoura, L.A.M.; D’Oca, M.G.M.; Banzato, T.P.; Monteiro, P.A.; Pilli, R.A.; de Carvalho, J.E.; Russowsky, D. Synthesis of Novel Perillyl-Dihydropyrimidinone Hybrids Designed for Antiproliferative Activity. Med. Chem. Commun. 2018, 9, 1553–1564. [Google Scholar] [CrossRef]
  10. Sharma, M.; Chauhan, K.; Shivahare, R.; Vishwakarma, P.; Suthar, M.K.; Sharma, A.; Gupta, S.; Saxena, J.K.; Lal, J.; Chandra, P.; et al. Discovery of a New Class of Natural Product-Inspired Quinazolinone Hybrid as Potent Antileishmanial agents. J. Med. Chem. 2013, 56, 4374–4392. [Google Scholar] [CrossRef]
  11. Camps, P.; Achab, R.E.; Morral, J.; Muñoz-Torrero, D.; Badia, A.; Baños, J.E.; Vivas, N.M.; Barril, X.; Orozco, M.; Luque, F.J. New Tacrine-Huperzine A Hybrids (Huprines): Highly Potent Tight-Binding Acetylcholinesterase Inhibitors of Interest for the Treatment of Alzheimer’s Disease. J. Med. Chem. 2000, 43, 4657–4666. [Google Scholar] [CrossRef]
  12. Singh, P.; Anand, A.; Kumar, V. Recent developments in biological activities of chalcones: A mini review. Eur. J. Med. Chem. 2014, 85, 758–777. [Google Scholar] [CrossRef] [PubMed]
  13. Zhuang, C.; Zhang, W.; Sheng, C.; Zhang, W.; Xing, C.; Miao, Z. Chalcone: A Privileged Structure in Medicinal Chemistry. Chem. Rev. 2017, 117, 7762–7810. [Google Scholar] [CrossRef] [PubMed]
  14. Rani, A.; Anand, A.; Kumar, K.; Kumar, V. Recent developments in biological aspects of chalcones: The odyssey continues. Expert Opin. Drug Discov. 2019, 14, 249–288. [Google Scholar] [CrossRef] [PubMed]
  15. Salum, L.B.; Altei, W.F.; Chiaradia, L.D.; Cordeiro, M.N.S.; Canevarolo, R.R.; Melo, C.P.S.; Winter, E.; Mattei, B.; Daghestani, H.N.; Santos-Silva, M.C.; et al. Cytotoxic 3,4,5-Trimethoxychalcones as Mitotic Arresters and Cell Migration Inhibitors. Eur. J. Med. Chem. 2013, 63, 501–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Shankaraiah, N.; Nekkanti, S.; Brahma, U.R.; Praveen Kumar, N.; Deshpande, N.; Prasanna, D.; Senwar, K.R.; Jaya Lakshmi, U. Synthesis of Different Heterocycles-Linked Chalcone Conjugates as Cytotoxic Agents and Tubulin Polymerization Inhibitors. Bioorgan. Med. Chem. 2017, 25, 4805–4816. [Google Scholar] [CrossRef]
  17. Wei, H.; Ruan, J.; Zhang, X. Coumarin-chalcone hybrids: Promising agents with diverse pharmacological properties. RSC Adv. 2016, 6, 10846–10860. [Google Scholar] [CrossRef]
  18. Mao, Z.; Zheng, X.; Qi, Y.; Zhang, M.; Huang, Y.; Wan, C.; Rao, G. Synthesis and biological evaluation of novel hybrid compounds between chalcone and piperazine as potential antitumor agents. RSC Adv. 2016, 6, 7723–7727. [Google Scholar] [CrossRef]
  19. Stanojković, T.; Marković, V.; Matić, I.Z.; Mladenović, M.P.; Petrovića, N.; Krivokuća, A.; Petković, M.; Joksović, M.D. Highly selective anthraquinone-chalcone hybrids as potential antileukemia agents. Bioorgan. Med. Chem. Lett. 2018, 28, 2593–2598. [Google Scholar] [CrossRef]
  20. Park, S.; Kim, E.H.; Kim, J.; Kim, S.H.; Kim, I. Biological evaluation of indolizine-chalcone hybrids as new anticancer agents. Eur. J. Med. Chem. 2018, 144, 435–443. [Google Scholar] [CrossRef]
  21. Thiriveedhi, A.; Nadh, R.V.; Srinivasu, N.; Kaushal, K. Novel Hybrid Molecules of Quinazoline Chalcone Derivatives: Synthesis and Study of In Vitro Cytotoxic Activities. Lett. Drug Des. Discov. 2018, 15, 757–765. [Google Scholar] [CrossRef]
  22. Mass, E.B.; Duarte, G.V.; Russowsky, D. The Quinazoline-Chalcone and Quinazolinone-Chalcone Hybrids: A Promising Combination for Biological Activity. Mini Rev. Med. Chem. 2021, 21, 186–203. [Google Scholar] [CrossRef] [PubMed]
  23. Jernei, T.; Duró, C.; Dembo, A.; Lajkó, E.; Takács, A.; Kohidai, L.; Schlosser, G.; Csámpai, A. Synthesis, Structure and In Vitro Cytotoxic Activity of Novel Cinchona-Chalcone Hybrids with 1,4-Disubstituted and 1,5-Disubstituted 1,2,3-Triazole Linkers. Molecules 2019, 24, 4077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Abbas, S.H.; El-Hafeez, A.A.A.; Shoman, M.E.; Montano, M.M.; Hassan, H.A. New quinoline-chalcone hybrids as anti-cancer agents: Design, synthesis, and evaluations of cytotoxicity and PI3K inhibitory activity. Bioorgan. Chem. 2019, 82, 360–377. [Google Scholar] [CrossRef] [PubMed]
  25. Chouiter, M.I.; Boulebd, H.; Pereira, D.M.; Valentao, P.; Andrade, P.B.; Belfaitah, A.; Silva, A.M.S. New chalcone-type compounds and 2-pyrazoline derivatives: Synthesis and caspase-dependent anticancer activity. Future Med. Chem. 2020, 12, 493–509. [Google Scholar] [CrossRef] [PubMed]
  26. Djemoui, A.; Naouri, A.; Ouahrani, M.R.; Djemoui, D.; Lahcene, S.; Lahrech, M.B.; Boukenna, L.; Albuquerque, H.M.T.; Saher, L.; Rocha, D.H.A.; et al. A step-by-step synthesis of triazole-benzimidazole-chalcone hybrids: Anticancer activity in human cells. J. Mol. Struct. 2020, 1204, 127487. [Google Scholar] [CrossRef]
  27. Bukhari, S.N.A.; Jasamai, M.; Jantan, I.; Ahmad, W. Review of Methods and Various Catalysts Used for Chalcone Synthesis. Mini Rev. Org. Chem. 2013, 10, 73–83. [Google Scholar] [CrossRef]
  28. Qin, H.L.; Zhang, Z.W.; Lekkala, R.; Alsulami, H.; Rakesh, K.P. Chalcone hybrids as privileged scaffolds in antimalarial drug discovery: A key review. Eur. J. Med. Chem. 2020, 193, 112215. [Google Scholar] [CrossRef]
  29. Gao, F.; Huang, G.; Xiao, J. Chalcone hybrids as potential anticancer agents: Current development, mechanism of action, and structure-activity relationship. Med. Res. Rev. 2020, 40, 2049–2084. [Google Scholar] [CrossRef]
  30. Hein, J.E.; Fokin, V.V. Copper-catalyzed azide–alkyne cycloaddition (CuAAC) and beyond: New reactivity of copper(I) acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. [Google Scholar] [CrossRef]
  31. Neumann, S.; Biewend, M.; Rana, S.; Binder, W.H. The CuAAC: Principles, homogeneous and heterogeneous catalysts, and novel developments and applications. Macromol. Rapid Commun. 2019, 41, 1900359. [Google Scholar] [CrossRef]
  32. Lal, K.; Yadav, P.; Kumar, A.; Kumar, A.; Paul, A.K. Design, synthesis, characterization, antimicrobial evaluation and molecular modeling studies of some dehydroacetic acid-chalcone-1,2,3-triazole hybrids. Bioorgan. Chem. 2018, 77, 236–244. [Google Scholar] [CrossRef] [PubMed]
  33. Singh, P.; Raj, R.; Kumar, V.; Mahajan, M.P.; Bedi, P.M.S.; Kaur, T.; Saxena, A.K. 1,2,3-Triazole tethered β-lactam-chalcone bifunctional hybrids: Synthesis and anticancer evaluation. Eur. J. Med. Chem. 2012, 47, 594–600. [Google Scholar] [CrossRef] [PubMed]
  34. Guantai, E.M.; Ncokazi, K.; Egan, T.J.; Gut, J.; Rosenthal, P.J.; Smith, P.J.; Chibale, K. Design, synthesis and in vitro antimalarial evaluation of triazole-linked chalcone and dienone hybrid compounds. Bioorgan. Med. Chem. 2010, 18, 8243–8256. [Google Scholar] [CrossRef] [PubMed]
  35. Mayer, T.U.; Kapoor, T.M.; Haggarty, S.J.; King, R.W.; Schreiber, S.L.; Mitchison, T.J. Small molecule inhibitor of mitotic spindle bipolarity identified in a phenotype-based screen. Science 1999, 286, 971–974. [Google Scholar] [CrossRef] [Green Version]
  36. Cochran, J.C.; Gatial, J.E.; Kapoor, T.M.; Gilbert, S.P. Monastrol inhibition of the mitotic kinesin Eg5. J. Biol. Chem. 2005, 280, 12658–12667. [Google Scholar] [CrossRef] [Green Version]
  37. Matos, L.H.S.; Masson, F.T.; Simeoni, L.A.; Homem-de-Mello, M. Biological Activity of Dihydropyrimidinone (DHPM) Derivatives: A Systematic Review. Eur. J. Med. Chem. 2018, 143, 1779–1789. [Google Scholar] [CrossRef]
  38. Kaur, R.; Chaudhary, S.; Kumar, K.; Gupta, M.K.; Rawal, R.K. Recent synthetic and medicinal perspectives of dihydropyrimidinones: A Review. Eur. J. Med. Chem. 2017, 132, 108–134. [Google Scholar] [CrossRef]
  39. de Fátima, Â.; Braga, T.C.; Neto, L.S.; Terra, B.S.; Oliveira, B.G.F.; da Silva, D.L.; Modolo, L.V. A Mini-Review on Biginelli adducts with notable pharmacological properties. J. Adv. Res. 2015, 6, 363–373. [Google Scholar] [CrossRef] [Green Version]
  40. Oliveira, F.S.; Oliveira, P.M.; Farias, L.M.; Brinkerhoff, R.C.; Sobrinho, R.C.M.A.; Treptow, T.M.; D’Oca, C.R.M.; Marinho, M.A.G.; Hort, M.A.; Horn, A.P.; et al. Synthesis and antitumoral activity of novel hybrid monastrol-fatty acids against glioma cells. Med. Chem. Commun. 2018, 9, 1282–1288. [Google Scholar] [CrossRef]
  41. Souza, V.P.; Santos, F.S.; Rodembusch, F.S.; Braga, C.B.; Ornelas, C.; Pilli, R.A.; Russowsky, D. Hybrid 3,4-dihydropyrimidin-2-(thi)ones as dual-functional bioactive molecules: Fluorescent probes and cytotoxic agents to cancer cells. New J. Chem. 2020, 44, 12440. [Google Scholar] [CrossRef]
  42. Malik, M.S.; Ahmed, S.A.; Althagafi, I.I.; Ansari, M.A.; Kama, A. Application of triazoles as bioisosteres and linkers in the development of microtubule targeting agents. RSC Med. Chem. 2020, 11, 327–348. [Google Scholar] [CrossRef] [PubMed]
  43. Hau, S.C.K.; Cheng, P.S.; Mak, T.C.W. Assembly of organosilver (I) frameworks with terminal ethynide and ethenyl groups on separate pendent arms attached to an aromatic ring. Polyhedron 2013, 52, 992–1008. [Google Scholar] [CrossRef]
  44. Chang, M.Y.; Tai, H.Y.; Chen, Y.L.; Hsu, R.T. Synthesis of 1,3-Diaryl-1H-Benzo[g]Indazoles. Tetrahedron 2012, 68, 7941–7948. [Google Scholar] [CrossRef]
  45. Lebed, P.S.; Kos, P.O.; Polovinko, V.V.; Tolmachev, A.A.; Vovk, M.V. Synthesis of new polyfunctional 5,6,7,8-tetrahydroimidazo-[1,5-c]pyrimidin-5-ones by the aza-Wittig reaction followed by intramolecular cyclization and 1,3-prototropic Shift. Russ. J. Org. Chem. 2009, 45, 921–927. [Google Scholar] [CrossRef]
  46. Chen, J.; Fu, X.; Zhou, L.; Zhang, J.; Qi, X.; Cao, X. A convergent route for the total synthesis of Malyngamides O, P, Q and R. J. Org. Chem. 2009, 74, 4149–4157. [Google Scholar] [CrossRef]
  47. Moro, A.V.; Ferreira, P.C.; Migowski, P.; Rodembusch, F.S.; Dupont, J.; Lüdtke, D.S. Synthesis and photophysical properties of fluorescent 2,1,3-benzothiadiazole-triazole-linked glycoconjugates: Selective chemosensors for Ni (II). Tetrahedron 2013, 69, 201–206. [Google Scholar] [CrossRef]
  48. Priti, K.; Nagarajan, A.; Uchil, P.D. Analysis of Cell Viability by the MTT Assay. Cold Spring Harb. Protoc. 2018, 6, 469–471. [Google Scholar]
  49. Afonso de Lima, C.; de Souza Bueno, I.L.; Nunes Siqueira Vasconcelos, S.; Sciani, J.M.; Ruiz, A.L.T.G.; Foglio, M.A.; Carvalho, J.E.D.; Barbarini Longato, G. Reversal of Ovarian Cancer Cell Lines Multidrug Resistance Phenotype by the Association of Apiole with Chemotherapies. Pharmaceuticals 2020, 13, 327. [Google Scholar] [CrossRef]
  50. Suffness, M.; Pezzuto, J.M. Methods in Plant Biochemistry: Assays for Bioactivity; Hostettmann, K., Ed.; Academic Press: London, UK, 1990; Volume 6, pp. 71–133. [Google Scholar]
  51. Zhou, W.; Pan, T.; Cui, H.; Zhao, Z.; Chu, P.K.; Yu, X.-F. Black Phosphorus: Bioactive Nanomaterials with Inherent and Selective Chemotherapeutic Effects. Angew. Chem. Int. Ed. 2019, 58, 769–774. [Google Scholar] [CrossRef]
  52. Brooks, S.C.; Locke, E.R.; Soule, H.D. Estrogen Receptor in a Human Cell Line (MCF-7) from Breast Carcinoma. J. Biol. Chem. 1973, 17, 6251–6253. [Google Scholar] [CrossRef]
  53. Penot, G.; Peron, C.L.; Merot, Y.; Grimaud-Fanouillère, E.; Ferrière, F.; Boujrad, B.; Kah, O.; Saligaut, C.; Ducouret, B.; Métivier, R.; et al. The Human Estrogen Receptor-α Isoform hERα46 Antagonizes the Proliferative Influence of hERα66 in MCF7 Breast Cancer Cells. Endocrinology 2005, 12, 5474–5484. [Google Scholar] [CrossRef] [PubMed]
  54. Mota, A.D.; Evangelista, A.F.; Macedo, T.; Oliveira, R.; Scapulatempo Neto, C.; Vieira, R.A.; Marques, M.M. Molecular characterization of breast cancer cell lines by clinical immunohistochemical markers. Oncol. Lett. 2017, 13, 4708–4712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Paterni, I.; Granchi, C.; Katzenellenbogen, J.A.; Minutolo, F. Estrogen receptors alpha (ERα) and beta (ERβ): Subtype-selective ligands and clinical potential. Steroids 2014, 90, 13–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Hans, R.H.; Guantai, E.M.; Lategan, C.; Smith, P.J.; Wan, B.; Franzblau, S.G.; Gut, J.; Rosenthal, P.J.; Chibale, K. Synthesis, Antimalarial and Antitubercular Activity of Acetylenic Chalcones. Bioorgan. Med. Chem. Lett. 2009, 20, 942–944. [Google Scholar] [CrossRef]
  57. Hussaini, S.M.A.; Yedla, P.; Babu, K.S.; Shaik, T.B.; Chityal, G.K.; Kamal, A. Synthesis and Biological Evaluation of 1,2,3-Triazole Tethered Pyrazoline and Chalcone Derivatives. Chem. Biol. Drug Des. 2016, 88, 97–109. [Google Scholar] [CrossRef] [PubMed]
  58. Laia, F.M.R.; Pinho, E.; Melo, T.M.V.D. Synthesis and reactivity of aziridines with internal dipolarophiles: An approach to 1,4-Dihydrochromeno-[4,3-b]pyrroles and 3-Methylenechromano-[4,3-b]pyrroles. Synthesis 2015, 47, 2781–2790. [Google Scholar]
  59. Singh, G.; Singh, J.; Mangat, S.S.; Arora, A. Synthetic approach towards ‘Click’ modified chalcone based organo-triethoxysilanes, UV-Vis Study. RSC Adv. 2014, 4, 60853–60865. [Google Scholar] [CrossRef]
  60. Niu, C.; Li, G.; Tuerxuntayi, A.; Aisa, H.A. Synthesis and bioactivity of new chalcone derivatives as potential tyrosinase activator based on the Click chemistry. Chin. J. Chem. 2015, 33, 486–494. [Google Scholar] [CrossRef]
  61. Sharghi, H.; Jokar, M. Al2O3/MeSO3H: A novel and recyclable catalyst for one-pot synthesis of 3,4-Dihydropyrimidinones or their sulfur derivatives in Biginelli condensation. Synth. Commun. 2009, 39, 958–979. [Google Scholar] [CrossRef] [Green Version]
  62. Mosmann, T. Rapid Colorimetric Assay for Cellular Growth and Survival: Application to Proliferation and Cytotoxicity Assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  63. Franco, Y.E.M.; Okubo, M.Y.; Torre, A.D.; Paiva, P.P.; Rosa, M.N.; Silva, V.A.O.; Reis, R.M.; Ruiz, A.L.T.G.; Imamura, P.M.; de Carvalho, J.E.; et al. Coronarin D Induces Apoptotic Cell Death and Cell Cycle Arrest in Human Glioblastoma Cell Line. Molecules 2019, 24, 4498. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of hybrid Chalcone compounds IIII.
Figure 1. Structures of hybrid Chalcone compounds IIII.
Ddc 01 00002 g001
Figure 2. Monastrol (IV) and bioactive DHPM hybrids VVII.
Figure 2. Monastrol (IV) and bioactive DHPM hybrids VVII.
Ddc 01 00002 g002
Scheme 1. Preparation of propargyloxy chalcones 4ae.
Scheme 1. Preparation of propargyloxy chalcones 4ae.
Ddc 01 00002 sch001
Scheme 2. Preparation of azido-dihydropyrimidinones 8a,b.
Scheme 2. Preparation of azido-dihydropyrimidinones 8a,b.
Ddc 01 00002 sch002
Scheme 3. Preparation of Chalcone-Dihydropyrimidinones 9aj.
Scheme 3. Preparation of Chalcone-Dihydropyrimidinones 9aj.
Ddc 01 00002 sch003
Figure 3. Molecular docking of hybrids to ERα receptor (6CHZ PDB code). (A) Hybrid 9d and antagonist H3B-9224 to show the entire protein and positioned molecules–the same pattern was observed for the other hybrids; (B) Detail of the hybrid 9d; (C) Hybrid 9g; (D) Hybrid 9h. The hybrids are shown in blue, the ERα receptor in green, and the antagonist is used as a model in magenta.
Figure 3. Molecular docking of hybrids to ERα receptor (6CHZ PDB code). (A) Hybrid 9d and antagonist H3B-9224 to show the entire protein and positioned molecules–the same pattern was observed for the other hybrids; (B) Detail of the hybrid 9d; (C) Hybrid 9g; (D) Hybrid 9h. The hybrids are shown in blue, the ERα receptor in green, and the antagonist is used as a model in magenta.
Ddc 01 00002 g003
Figure 4. Molecular docking of hybrids to ERβ receptor (3OLS PDB code). (A) Hybrid 9d, (B) Hybrid 9g, and (C) Hybrid 9h, in blue. The Erβ receptor (3OLS PDB code) is in green, and the agonist estradiol is used as a model in magenta.
Figure 4. Molecular docking of hybrids to ERβ receptor (3OLS PDB code). (A) Hybrid 9d, (B) Hybrid 9g, and (C) Hybrid 9h, in blue. The Erβ receptor (3OLS PDB code) is in green, and the agonist estradiol is used as a model in magenta.
Ddc 01 00002 g004
Table 1. Synthesis of hybrids chalcone-DHPM 9aj.
Table 1. Synthesis of hybrids chalcone-DHPM 9aj.
EntryHybrid CompoundsH-Triazole (ppm)Yield (%) a
19aDdc 01 00002 i0019.0075
29bDdc 01 00002 i0028.3570
39cDdc 01 00002 i0038.4378
49dDdc 01 00002 i0048.4890
59eDdc 01 00002 i0058.3975
69fDdc 01 00002 i0069.3972
79gDdc 01 00002 i0078.4270
89hDdc 01 00002 i0088.5662
99iDdc 01 00002 i0098.4990
109jDdc 01 00002 i0108.4385
a Yields correspond to the purified compounds.
Table 2. IC50 values of hybrid compounds 9aj against different cell lines a.
Table 2. IC50 values of hybrid compounds 9aj against different cell lines a.
HybridU-251MCF-7NCI/ADR-RESOVCAR-3HT-29HaCaT
9a27.0 ± 11.06.2 ± 3.46.4 ± 3.49.9 ± 3.518.6 ± 8.76.7 ± 3.0
9b26.1 ± 8.25.3 ± 2.714.5 ± 5.59.4 ± 3.443.8 ± 12.822.0 ± 3.9
9c65.1 ± 11.011.5 ± 5.154.8 ± 9.670.6 ± 12.1134.1 ± 33.722.0 ± 3.9
9d74.6 ± 17.54.7 ± 2.589.8 ± 12.1109.7 ± 31.8109.7 ± 46.517.5 ± 2.3
9e24.6 ± 6.612.3 ± 3.426.8 ± 7.215.3 ± 3.441.1 ± 7.912.0 ± 2.2
9f10.9 ± 4.17.5 ± 2.740.4 ± 11.212.5 ± 4.725.1 ± 12.56.4 ± 3.0
9g13.3 ± 3.84.9 ± 1.439.6 ± 7.917.9 ± 5.893.9 ± 29.815.1 ± 4.1
9h70.4 ± 9.55.8 ± 2.6>153.036.4 ± 7.2>153.0100.7 ± 13.0
9i104.1 ± 33.25.3 ± 2.5>146.323.5 ± 6.0103.4 ± 18.134.8 ± 4.8
9j128.7 ± 35.514.6 ± 4.5>146.388.9 ± 11.0>146.369.3 ± 7.9
DOXO>18.40.4 ± 0.022.6 ± 3.911.6 ± 5.1>18.41.6 ± 0.0
a Values of IC50 are expressed in µM, triplicate. U-251 (brain, glioblastoma), MCF-7 (breast, adenocarcinoma), NCI/ADR-RES (ovary, multidrug-resistant adenocarcinoma), OVCAR-3 (ovary, adenocarcinoma), HT-29 (colorectal, adenocarcinoma) and HaCaT (normal human keratinocytes). Doxo: doxorubicin.
Table 3. Comparative IC50 (µM) and selectivity index (SI) values.
Table 3. Comparative IC50 (µM) and selectivity index (SI) values.
Cell LinesChalconeDHPMHybridSelectivity Index (SI)
4d8a9d
(4d + 8a)
9d
U-251251.51>33274.620.23
MCF-7>342>3324.723.71
NCI/ADR-RES>342>33289.790.19
HaCaT97.05>33217.52-
4b8b9g
(4b + 8b)
9g
U-25124.47>25513.301.14
MCF-780.33>2554.893.10
NCI/ADR-RES>381>25539.620.38
HaCaT23.33>25515.14-
4c8b9h
(4c + 8b)
9h
U-251126.50>25570.371.43
MCF-7>381>2555.8117.33
NCI/ADR-RES>381>255>153-
HaCaT37.59>255100.7-
Table 4. Molecular interactions between hybrids and the amino acid residues in ER-α (6CHZ) obtained by molecular docking a.
Table 4. Molecular interactions between hybrids and the amino acid residues in ER-α (6CHZ) obtained by molecular docking a.
HybridBinding Energy (kcal/mol)Binding InteractionBond Lenght (Å)
9d−10.32Leu3466.96
Thr347a3.93
Asp351a3.29
Glu353a8.27
9g−12.02Leu3466.33
Thr347a4.05
Asp351a3.68
Asn5322.36
Leu5362.50
9h−7.52Leu3464.86
Thr347a5.35
Asp351a4.27
a Binding pocket: Thr347, Asp351, Glu353 for the endogenous agonist.
Table 5. Molecular interactions between hybrids and the amino acid residues in ER-β1 (3OLS) obtained by molecular docking a.
Table 5. Molecular interactions between hybrids and the amino acid residues in ER-β1 (3OLS) obtained by molecular docking a.
HybridBinding Energy (kcal/mol)Binding InteractionBond Lenght (Å)
9d−9.38Cys5302.38
Ser5379.88
Asp3514.46
9g−7.85Cys5302.59
Ser5377.26
Leu5362.36
9h−8.51Cys5304.24
Ser5377.01
Leu5367.62
a Binding pocket: Glu305, Arg346, Phe356 for the endogenous agonist.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mass, E.B.; de Lima, C.A.; D’Oca, M.G.M.; Sciani, J.M.; Longato, G.B.; Russowsky, D. Synthesis, Selective Cytotoxic Activity against Human Breast Cancer MCF7 Cell Line and Molecular Docking of Some Chalcone-Dihydropyrimidone Hybrids. Drugs Drug Candidates 2022, 1, 3-21. https://doi.org/10.3390/ddc1010002

AMA Style

Mass EB, de Lima CA, D’Oca MGM, Sciani JM, Longato GB, Russowsky D. Synthesis, Selective Cytotoxic Activity against Human Breast Cancer MCF7 Cell Line and Molecular Docking of Some Chalcone-Dihydropyrimidone Hybrids. Drugs and Drug Candidates. 2022; 1(1):3-21. https://doi.org/10.3390/ddc1010002

Chicago/Turabian Style

Mass, Eduardo B., Carolina A. de Lima, Marcelo G. M. D’Oca, Juliana M. Sciani, Giovanna B. Longato, and Dennis Russowsky. 2022. "Synthesis, Selective Cytotoxic Activity against Human Breast Cancer MCF7 Cell Line and Molecular Docking of Some Chalcone-Dihydropyrimidone Hybrids" Drugs and Drug Candidates 1, no. 1: 3-21. https://doi.org/10.3390/ddc1010002

Article Metrics

Back to TopTop