Next Article in Journal
Carex meyeriana Kunth Extract Is a Novel Natural Drug against Candida albicans
Previous Article in Journal
Vitamin C Supplementation in the Treatment of Autoimmune and Onco-Hematological Diseases: From Prophylaxis to Adjuvant Therapy
Previous Article in Special Issue
Evaluation of Antitumor Activity of Xanthones Conjugated with Amino Acids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Anticancer Activity, and Docking Studies of Novel Hydroquinone-Chalcone-Pyrazoline Hybrid Derivatives

Instituto de Química, Facultad de Ciencias, Pontificia Universidad Católica de Valparaíso, Valparaíso 23732223, Chile
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(13), 7281; https://doi.org/10.3390/ijms25137281
Submission received: 8 June 2024 / Revised: 20 June 2024 / Accepted: 30 June 2024 / Published: 2 July 2024
(This article belongs to the Special Issue Bioactive Compounds in Cancers)

Abstract

:
A novel series of antitumor hybrids was synthesized using 1,4-benzohydroquinone and chalcone, furane, or pyrazoline scaffolds. This were achieved through isosteric substitution of the aryl group of the chalcone β-carbon with the furanyl moiety and structural modification of the α,β-unsaturated carbonyl system. The potential antitumor activity of these hybrids was evaluated in vivo on MCF-7 breast adenocarcinoma and HT-29 colorectal carcinoma cells, demonstrating cytotoxic activity with IC50 values ranging from 28.8 to 124.6 µM. The incorporation of furan and pyrazoline groups significantly enhanced antiproliferative properties compared to their analogues and precursors (VIIX), which were inactive against both neoplastic cell lines. Compounds 4, 5, and 6 exhibited enhanced cytotoxicity against both cell lines, whereas compound 8 showed higher cytotoxic activity against HT-29 cells. Molecular docking studies revealed superior free-energy values (ΔGbin) for carcinogenic pathway-involved kinase proteins, with our in silico data suggesting that these derivatives could be promising chemotherapeutic agents targeting kinase pathways. Among all the synthesized PIBHQ compounds, derivatives 7 and 8 exhibited the best drug-likeness properties, with values of 0.53 and 0.83, respectively. ADME results collectively suggest that most of these compounds hold promise as potential candidates for preclinical assays.

1. Introduction

Cancer results from uncontrolled cell division caused by deregulation of the biological mechanisms controlling the cell cycle [1]. The sustained increases in cancer incidence and mortality over the last decade necessitate the discovery of new compounds with superior anticancer properties compared to current chemotherapeutic drugs [2].
Hydroquinone represents a simple phenolic base structure found as secondary metabolites in various natural products and serves as a privileged scaffold in bioorganic and medicinal chemistry [3,4,5,6]. These compounds exhibit a wide range of biological activities, including antineoplastic, anti-inflammatory, antileishmanial, antimicrobial, and antioxidant properties [4,7,8,9]. Several hydroquinones have demonstrated antitumor activity against breast, colon, lung, and cervical cancers, among others [7,10,11]. Their antineoplastic properties are primarily attributed to their ability to generate reactive oxygen species (ROS) and induce significant DNA damage [12,13]. Specifically, alkyl hydroquinones have been implicated in antiproliferative processes through various mechanisms, such as downregulation of p53, inhibition of protein kinases, tubulin polymerization, and cyclooxygenases [8,9,14,15,16].
Chalcones are structures of both natural and synthetic origin that exhibit various pharmacological activities, including antineoplastic, antioxidant, anti-inflammatory, antiviral, antimalarial, antimicrobial, analgesic, and antitubercular properties [17,18,19,20]. Studies on their anticancer properties have shown that they interact with various target proteins involved in proliferative processes. One of the most recognized antineoplastic mechanisms of these derivatives is the interruption of the cell cycle through microtubule disruption by binding to tubulins, preventing their polymerization. This disruption in mitotic spindle assembly and cytoskeletal function results in cell cycle arrest in the G2/M phase [21]. Additionally, chalcones can activate p53 protein; inhibit topoisomerase activity; halt angiogenesis by inhibiting vascular endothelial growth factor (VEGF); and inhibit various enzymes such as androgen receptor, kinases, mesenchymal–epithelial transition factor (MET), cyclooxygenase 2 (COX-2), histone deacetylases (HDACs), nuclear factor kappa B (NF-κB), and epidermal growth factor receptor (EGFR). These biological processes collectively induce apoptosis [22,23,24].
Chalcones containing aromatic heterocyclic groups in their structure enhance the aforementioned biological activities [25,26]. In particular, the furan ring in chalcone structures has shown significant anticancer properties [20,21,27,28]. Studies have indicated that the bioisosteric exchange of the aryl group for this aromatic heterocycle in chalcones leads to the synthesis of potent inhibitors of protein tyrosine phosphatase 1B (PTP1B) [29].
Pyrazoline is a five-membered heterocycle containing two nitrogen atoms and holds great pharmacological interest. This versatile scaffold is used to develop new anticancer agents, and many clinically used anticancer drugs contain this group [30]. Specifically, pyrazoline derivatives have been involved in antineoplastic processes through inhibition of VEGFR-2, EGFR, tubulin polymerization deregulation, and inhibition of topoisomerase II. These processes inhibit cell division, DNA synthesis, or cell migration [31,32,33,34].
Molecular hybridization is a strategy in medicinal chemistry to discover new compounds with pharmacological properties by combining two or more structural units with recognized biological activities into a single structure [22,35,36]. In previous studies, we synthesized a series of hybrid derivatives of naphthoquinones conjugated with isoxazole [37,38] and pyrazole [39,40], as well as chalcones fused to naphthoquinone/hydroquinone systems with promising antineoplastic properties [41]. In this work, we rationally designed and synthesized a new series of chalcone–hydroquinone hybrids, including structural modifications of compounds previously prepared by our research group [41]. We performed an isosteric replacement of the aryl group of the chalcone β-carbon with the furanyl heterocycle. Additionally, by modifying the α,β-unsaturated carbonyl system, we synthesized novel pyrazoline-1,4-benzohydroquinone (PIBHQ) hybrids (Figure 1). Finally, we assessed the in vitro antineoplastic activity of these hybrids on MCF-7 and HT-29 cancer cell lines and conducted in silico molecular docking studies to identify potential biological targets through virtual screening. This included various cancer-related proteins overexpressed in breast and colon cancers to evaluate the antitumor potential and explore the possible action mechanisms of these new hybrid compounds.

2. Results and Discussion

2.1. Chemistry

As shown in Scheme 1, the synthesis of compounds 3 and 4 was derived from the precursor 1-{1-hydroxy-6-(4-methylpent-3-en-1-yl)-4-[(tetrahydro-2H-pyran-2-yl)oxy]-5,8-dihydronaphthalen-2-yl}ethanone naphthohydroquinone (2). Initially, the phenolic -OH group attached to C-4 of compound 1 was protected using 3,4-dihydro-2H-pyran (DHP) in the presence of pyridinium p-toluenesulfonate (PPTS) as a catalyst according to our previously described procedure [41]. This step was necessary to eliminate the acidic properties of the phenolic group.
The new hybrid, chalcone/1,4-benzohydroquinone (CBHQ) (4), was synthesized via Claisen–Schmidt condensation of precursor 2 with 2-furfuraldehyde, yielding derivative 3 with a moderate yield of 63%. Subsequent acid hydrolysis using 4-toluenesulfonic acid monohydrate (PTSA) allowed for the high-yield (92%) production of CBHQ 4.
The structures of all new compounds were determined by IR, 1H NMR (400 MHz), 13C NMR (100 MHz), and elemental analysis. The infrared (IR) spectra of derivatives 3 and 4 exhibited characteristic absorption bands of the stretching vibration (stv) of the phenolic O-H bonds in the range of 3374–3481 cm−1, in addition to the stv bands of the C=O bonds at 1634–1639 cm−1. The stv band of C=C bond of chalcones was observed in the range 1581–1583 cm−1. Moreover, the 1H NMR spectrum of compound 3 showed signals corresponding to the protons of the α,β-unsaturated fragment characteristic of H-16 and H-17 chalcones at δ 7.50 and δ 7.68 ppm, respectively. The analogous protons H-16 and H-17 of CBHQ 4 were observed at δ 7.45 and δ 7.65 ppm, respectively, with both protons of the unsaturated system showing a coupling constant (J) of 15.1 Hz. Phenolic protons 4-OH and 1-OH appeared as singlets at δ 4.80 and 13.00 ppm, respectively. This marked difference in chemical shift is due to the intramolecular interaction of the 1-OH proton with the carbonyl group of the chalcone fragment. Additionally, the signals for protons H-20, H-19, and H-21 of the furan heterocycle were observed at δ 6.54, 6.74, and 7.54 ppm, respectively.
In the 13C NMR spectrum of CBHQ 4, signals attributed to all carbons of the α,β-unsaturated carbonyl fragment characteristic of chalcones were observed. Particularly, the carbonyl at C-15 absorbed at δ 192.5 ppm, while the signals of carbons C-20, C-19, C-21, and C-18 of the furan heterocycle appeared at δ 112.8, 116.8, 145.3, and 151.6 ppm, respectively. The IR, 1H NMR, and 13C NMR spectra of compounds 3 and 4 are presented in Figures S1–S6.
As shown in Scheme 2, the synthesized CBHQ 4 facilitated the production of new compounds 5 and 6. For the synthesis of CBHQ 5, acetylation of the phenolic groups of CBHQ 4 was performed using acetic anhydride in pyridine, achieving a high yield of 96%. Additionally, the reactivity of the α, β-unsaturated carbonyl system of CBHQ 4 was exploited to carry out a conjugated addition with hydrazine, resulting in the hybrid PIBHQ 6 containing the pyrazoline heterocycle. This synthesis utilized excess hydrazine monohydrate in ethanol under reflux, yielding a product with a 71% yield (Scheme 2).
The infrared (IR) spectrum of acetylated CBHQ 5 showed an stv absorption band of the C=O of the acetyl groups at 1766 cm−1. The 1H NMR spectrum exhibited a multiplet signal corresponding to the protons of acetyl groups H-2′ and H-4′ at δ 2.26 ppm, while the signals of the H-16 and H-17 protons appeared as doublets at δ 7.10 and δ 7.43 ppm, respectively, with coupling constants (J) of 15.5 Hz. In the 13C spectrum, the signals corresponding to methyls C-2′ and C-4′ of the acetyl groups appeared at δ 20.8 and 20.9 ppm, respectively, while intense signals of both carbonyls of the C-3′ and C-1′ acetyl groups were observed at δ 169.0 and 169.1 ppm, respectively.
Regarding the structural characterization of PIBHQ 6, the IR spectrum displayed an stv band at 3328 cm−1 of the phenolic O-H bond and a N-H bond of pyrazoline, while the stv band of the C=N bond was observed at 3328 cm−1. The 1H NMR spectrum showed the signals corresponding to protons H-16, H-17, H-19, H-20, and H-21 at δ 3.29, 4.88, 6.23, 6.32, and 7.37 ppm, respectively. In the 13C NMR spectrum, signals C-16, C-17, C-19, C-20, and C-21 were observed at δ 37.9, 55.5, 106.4, 110.3, and 142.5 ppm, respectively. The C-15 signal of the pyrazoline heterocycle was also observed at a higher chemical shift to δ 155.1 ppm. The IR, 1H NMR, and 13C NMR spectra of compounds 5 and 6 are shown in Figures S7–S12.
Utilizing the structural similarity of compounds VIIX synthesized in previous studies [41] with the hybrid CBHQ 4, we proposed the synthesis of new hybrid molecules of pyrazoline-type PIBHQs (710). This was achieved through a cyclocondensation reaction of the α,β-unsaturated carbonyl fragment of precursors VIIX with hydrazine monohydrate under reflux heating in ethanol as a solvent (Scheme 3). PIBHQs 710 were obtained with moderate yields (62–75%), with higher yields observed for the 2,4-dichloro substituents in the phenyl group.
For the 1H NMR characterization of PIBHQs 710, the elucidation of the chemical skeleton of the pyrazolines focused on analyzing the signals corresponding to the ABX hydrogen system of the heterocyclic ring. The 1H NMR spectra showed the signal of the H-16A proton in the range δ 2.84–3.06 ppm, while the H-16B proton signal appeared between δ 3.36 and 3.61 ppm. The H-17X proton signal was observed in the range of δ 4.79–5.18 ppm. Generally, the magnetically non-equivalent diastereotopic protons of the methylene group H-16A and H-16B couple with the H-17X proton displayed doublet–doublet signals with geminal and vicinal coupling constants (JAB = 16.3–16.7 Hz, JAX = 8.1–9.4 Hz and JBX = 10.2–10.7 Hz). The IR, 1H NMR, and 13C NMR spectra of compounds 710 are shown in Figures S13–S24.

2.2. In Vitro Cytotoxicity Assays

The antiproliferative activity of the new compounds (410) assessed on MCF-7 and HT-29 cancer cell lines showed a cytotoxic effect due to the conjugation of chalcone and pyrazoline pharmacophores with the hydroquinone derivatives, a we previously evaluated in these cell cultures, which are among the most common types of cancer, as previously evaluated on cancer cell lines. The results are shown in Table 1 and are expressed as the half-maximal inhibitory concentration (IC50 ± standard deviation) of cell proliferation for each evaluated compound, along with the calculated −log10 of that value (pIC50).
As presented in Table 1, CBHQs 4 and 5 and PIBHQ 6 hybrids containing a furan ring exhibited a notable cytotoxic effect against both cancer cell lines. For MCF-7, the pIC50 values exceeded 4.46, while for HT-29, the pIC50 values were >4.42. CBHQ 5 demonstrated the highest cytotoxicity, with pIC50 values of 4.73 and 4.66 for HT-29 and MCF-7, respectively. These results indicate that diacetylation in CBHQ hybrids enhances cytotoxicity against both neoplastic cell lines, consistent with previously reported studies [41].
Cytotoxic activity of CBHQ hybrids VIIX from our previous studies was compared with that of the analogous compounds synthesized in this work, which involved the isosteric substitution of the aryl ring with the furanyl heterocycle in the chalcone moiety. The results showed that CBHQ 4 exhibited significant cytotoxicity against MCF-7 and HT-29 cell lines, with pIC50 values of 4.54 and 4.47, respectively, notably enhancing the antiproliferative properties compared to their VIIX analogues, which were inactive against both tested neoplastic cell lines [41].
PIBHQ derivative 6 generally showed better cytotoxicity values compared to analogous hybrids 710, increasing cytotoxic potency against both tumoral cell lines by up to three times (Figure 2). This underscores the importance of the furanyl heterocycle in this series of hybrids. The exception was compound 8, which exhibited a significant pIC50 value of 4.51 against the HT-29 cell line. Interestingly, the presence of 2,4-dichloro substituents in hybrid 10 resulted in limited antiproliferative effects, with pIC50 values lower than 3.52 against both cancer cell lines, similar to previously reported results [41].
Regarding structure–activity relationship (SAR) studies, the isosteric substitution of the benzene ring with the furan heterocycle in the chemical skeletons of chalcone/1,4-benzohydroquinone hybrids was found to be crucial for obtaining cytotoxic molecules. The respective analogues, CBHQs VIIX, were inactive against both studied cell lines [41]. Additionally, the incorporation of the pyrazoline heterocycle into the chemical structures of CBHQs VIIX (Figure 3) was significant in producing new cytotoxic molecules, as evidenced by PIBHQs 79, which showed pIC50 values between 3.90 and 4.51, compared to their inactive precursor chalcones (VIIX) against MCF-7 and HT-29 cell lines. These findings motivate us to continue develo** the designs of new bioactive compounds, leveraging the reactivity of chalcones to incorporate new heterocyclic bioisosteric entities, aiming to obtain new candidates with greater pharmacological potential as anticancer agents.

2.3. In Silico Virtual Screening for Potential Antineoplastic Targets of CBHQs and PIBHQs

To obtain information about the potential anticancer mechanisms of the synthesized compounds, we performed molecular docking studies to identify possible biological targets in various cancer-related proteins and calculated their corresponding ∆Gbin values. The searches for the possible binding sites of these cytotoxic derivatives were conducted through virtual screening with proteins involved in antineoplastic processes. Among these targets, we selected proteins overexpressed in breast cancer and colon cancer, including the mesenchymal–epithelial transition factor (c-MET), receptor of the epidermal growth factor (EGFR), fibroblast growth factor receptor 2 (FGFR-2), epidermal growth factor receptor 2 (HER2), tropomyosin receptor kinase A (TRKA), tyrosine protein kinase (TPK), mitogen-activated protein kinases (MAPK1, ERK2, and MEK1), dihydrofolate reductase (DHFR), estrogen receptors (ER), vascular endothelial growth factor receptor 2 (VEGFR-2), and tubulin (TUB), among others [42,43,44,45,46].
The virtual screening performed with the selected proteins (Table S1) revealed that the cytotoxic derivatives present more stable conformations, particularly with protein kinases, with binding energies averaging greater than −8.42 kcal/mol. As shown in Table 2, the derivatives primarily bind strongly to the epidermal growth factor receptor (EGFR), with ∆Gbin values ranging from −11.4 to −8.5 kcal/mol (average, −10.68 kcal/mol). This is followed by protein kinases, with the best bindings to MEK1 having energies between −10.7 and −8.5 kcal/mol (average, −9.95 kcal/mol); CK4, ranging from −10.5 to −9.3 kcal/mol (average, −9.87 kcal/mol); and TPK, ranging from −10.6 to −8.5 kcal/mol (average, −9.72 kcal/mol). Interestingly, PIBHQs 69 exhibited higher affinity for all the selected protein kinases (Table 2) compared to CBHQ hybrids 4 and 5, demonstrating the importance of the pyrazoline heterocycle for protein binding. Additionally, in vitro activity studies have shown that pyrazoline derivatives can act as EGFR kinase inhibitors [47,48,49,50].
As detailed in Table 2, the most favorable ∆Gbin values against EGFR and MEK1 were demonstrated by derivative 7 (−11.4 and −10.7 kcal/mol, respectively). The highest ∆Gbin values against CK4, TPK, and TRKA were exhibited by derivative 9 (−10.5, −10.6, and −10.1 kcal/mol, respectively), while compound 6 showed the highest affinity against TPK (−10.3 kcal/mol). Interestingly, PIBHQ hybrids 69 showed a slight increase in affinity for most protein kinases, with ∆Gbin values exceeding those of CBHQs 4 and 5, despite their greater cytotoxicity. Most of the cytotoxic hybrids exhibited better binding affinities to the studied proteins, with higher ∆Gbin values compared to the anticancer drugs used as references (anlotinib, crizotinib, and erlotinib), which act as strong kinase inhibitors [51,52,53,54,55,56,57,58,59,60]. Notably, most of the compounds demonstrated superior ΔGbin values compared to these reference antiproliferative drugs (Table 2).
Despite PIBHQ 10 not exhibiting cytotoxicity against the tested cancer cell lines (Table 1), it demonstrated favorable binding energies within the active sites of kinase proteins, as presented in Table S2. This observation can be attributed to robust hydrogen bond interactions between the free hydroxyl groups of the hydroquinone system and oxygen- or nitrogen-containing groups within the proteins, as well as van der Waals interactions between the chloride of the phenyl group and residues, as illustrated in Figures S25–S27.

2.4. Binding Sites and Docking of Synthesized Cytotoxic Hybrids in EGFR, MEK1, CK4, and TPK Targets

A more detailed analysis comparing the binding energies of PIBHQs with respect to their CBHQ precursors against EGFR, MEK1, CK4, and TPK (Table S2) indicates that most compounds show an increase in affinities against protein kinases, with ∆Gbin average values of less than -8.50 kcal/mol. The most favorable ∆Gbin values were shown against EGFR and MEK1, with average ∆Gbin energies of −10.68 and −9.95 kcal/mol, respectively, followed by c-MET and TRKA (−9.82 and −9.72 kcal/mol, respectively). These findings suggest that these hybrids might serve as potent inhibitors of EGFR, MEK1, and c-MET, all of which are overexpressed in certain types of cancer, including human breast and colorectal cancer [42,61,62,63]. Furthermore, PIBHQs 69 showed higher affinity for all the selected protein kinases (Table 2) compared to CBHQ hybrids 4 and 5. These findings reaffirm that the presence of the pyrazoline cycle in hybrids 79 is crucial for increased affinity to the selected protein kinases compared to the precursor chalcones (VIIIX).
Detailed configurations of the binding sites, along with the amino acids involved in the docking of synthesized cytotoxic hybrids and their corresponding ΔGbin values for EGFR, MEK1, and c-MET, are presented in Table 3 and depicted in 2D maps in Figure 4. Additionally, 3D docking complexes of c-MET with 69 are illustrated in Figure 4. Complementary 2D maps for complexes involving 4 and 5 can be found in Figure S17, and binding site interactions of the synthesized cytotoxic hybrids with amino acids of TPK, TRKA, and CK4 are outlined in Table S2.
Overall, as demonstrated in Table 2 and Table 3, PBHQ derivative 7 exhibited superior binding affinity for biological targets, with the best average ΔGbin values for EGFR and MEK1, due to the presence of the C1 and C4 hydroxyl groups of the hydroquinone ring. These groups interacted with amino acid residues through hydrogen bonding. For instance, the hydroxyl groups of the hydroquinone ring in derivatives 7 and 8 interacted with Ala743 of EGFR (Figure 4). Specifically, PBHQs 6, 7, 8, and 9 displayed excellent binding affinities for c-MET, with ΔGbin values of −11.3, −11.4, −11.0, and −11.2 kcal/mol, respectively. Peptide sequences surrounding the PIBHQs revealed consistent docking in the same region of the enzyme, defined by residues Ala743 and Met790. All the compounds, including 4 and 5, engaged in hydrogen bonding with EGFR residues, as well as various other interactions, including van der Waals, Pi–anion, Pi–sigma, Pi–Pi stacked, and Pi–alkyl interactions.
According to hydrogen bonds, residues Ala743, Lys745, and Leu788 were most commonly involved in interactions with EGFR, serving as hydrogen bond donors toward hydroxyl groups from the hydroquinone moieties of 4, 6, 7, and 8; the pyrazoline moiety of 8; or the acetyl group of 5 (Figure 5 and Figure S25). In the case of MEK1, residues Val211 and Ser212 interacted with hydroxyl groups from the hydroquinone of 4 and 7, respectively, while Lys97 interacted with the nitrogen from the pyrazoline moieties of 8 and 9 through hydrogen bonds (Figure S26). c-MET exhibited interactions with residues such as Arg1227, which primarily engaged in hydrogen bonds with hydroxyl groups from the hydroquinones of 4 and 9, as well as Asn1167, which interacted with the furan groups of 4 and 5, and Asn1167 and Arg1227 with the acetyl group of 5 (Figure S27). Tyr1234 and Arg1166 interacted with the oxygen of the methoxy group of 8.
Aromatic interactions, similar to hydrogen bonds, play a crucial role in ligand–protein interfaces. Many contemporary ligand docking programs implicitly account for aromatic stacking through van der Waals and Coulombic potentials [41,64]. Residues Met766, Met790, and Cys797 of EGFR were notably involved in these interactions, engaging in π–sulfur interactions with the aromatic rings of hydroquinone systems in 49 b (Figure 4 and Figure S25). Additionally, the sulfur of Met790 interacts with the rings of the hydroquinone in 4 and 5, in addition to Cys797 and Met766 interacting with the furan group in 4 and 5, respectively, through π–sulfur interactions (Figure S25).
In the case of MEK1, Phe209 was the primary residue involved in aromatic interactions, participating in π–π stacking with the furan group in 5 and the aromatic ring of the phenyl group in 9 and 10. The sulfur of Met143 interacted with the furan group in 5 and the aromatic ring of the phenyl group in 8 and 10 through π–sulfur interactions. Leu215 and Met219 interacted through π–alkyl interactions with the aromatic rings of the quinone moiety in 5–9, while Asp208 interacted through π–anion interactions with the aromatic ring of the pyrazoline heterocycle in 9 and 10 (Figure S26). Lastly, Phe1089 and Phe1223 played a prominent role in c-MET interactions, engaging in π–π stacking with the aromatic ring of the pyrazoline heterocycle in 6, 9, and 10. Arg1227 interacted through π–alkyl interactions with the aromatic rings of the quinone moiety in 4 and 5, while Val1092 interacted through π–sigma interactions with the aromatic rings of the pyrazoline heterocycle in 7 and 9. Asp1164 interacted through π–anion interactions with the hydroquinone ring in 9 and 10 (Figure S27).
To validate the binding sites of the synthesized cytotoxic hybrids within the kinases, we conducted a comparative analysis of PIBHQ 7 complexes with those of known kinase ligands. The results revealed that the binding regions of PIBHQ 7, indeed, overlap with the catalytic sites of the target enzymes, sharing a common set of contacts with the respective inhibitors (Table 4). Notably, the active site residues involved in these interactions include Leu718, Val726, Ala743, Lys745, Met766, Cys775, Arg776, Leu777, Leu788, Met790, Met793, Gly796, Leu844, Thr854, Asp855, Phe856, and Leu858 for EGFR; Lys97, Leu115, Leu118, Val127, Phe129, Ile141, Met143, Asp190, Cys207, Asp208, Phe209, Gly210, Val211, Ser212, Leu215, Ile216, and Met219 for MEK1; and Phe1089, Val1092, Lys1110, Leu1157, Met1211, Phe1223, and Ala1226 for c-MET. These residues served as common contact points for PIBHQ 7 and ligands 1, 2, and 3 in all three enzymes.
We compared the binding site and affinity of PIBHQ 7 with anticancer drugs used as inhibitors of EGFR, MEK1, and c-MET [54,55,57]. Notably, the energetic aspects of these interactions favored PIBHQ 7 in comparison to erlotinib, with a favorable energy difference of 1.5 kcal/mol for c-MET and 2.8 kcal/mol for EGFR. Moreover, 7 exhibited better in silico affinity than anlotinib for MEK1 and crizotinib for MEK1, with favorable energy differences of 1.0 kcal/mol and 0.3 kcal/mol, respectively. Importantly, the aromatic ring within the pyrazoline moiety, as well as the hydroxyl groups from the hydroquinone moiety of PIBHQs, plays a pivotal role in these interactions, directly contributing to the overlap with the ligands at the catalytic sites of the enzymes (Figure 6). This crucial involvement of the pyrazoline moiety is consistently observed in the case of EGFR, as well as MEK1 and c-MET (Figures S26 and S27).
Furthermore, it is possible to speculate that CBHQ 5 may act as a prodrug. It could undergo hydrolysis through deacetylation within the cell, catalyzed by a “deacetylase” enzyme, releasing the molecule in the form of CBHQ 4. Subsequently, these benzohydroquinone compounds might exhibit an inhibitory effect on cancer-related kinases. This assumption is supported by their favorable binding energies in the active site of the kinase domain of EGFR, as detailed in Figure S25 and Table 4.

2.5. In Silico Drug Likeness, Toxicity Risks, and ADME Predictions

Drug-likeness scores for compounds 49 were computed using the DataWarrior algorithm, and the results are presented in Table 5. Notably, derivatives 7 and 8 stand out as the only compounds with positive drug-likeness values of 0.53 and 0.83, respectively. This significant finding suggests that compounds 7 and 8 could be promising lead candidates for further investigation. It is noteworthy that compounds 7 and 8 incorporate essential structural elements, such as the hydroquinone and pyrazoline fragments, which are commonly found in approved drugs. Additionally, compound 8 features a phenyl fragment with a methoxy group at position 3. These substituents are known to contribute significantly to the enhancement of the antineoplastic cytotoxicity of potential anticancer agents.
In terms of toxicity risks, CBHQs 4 and 5 are likely to exhibit a high level of mutagenic risk, whereas PIBHQs 69 are expected to have no adverse effects (Table 5). The high irritant risk associated with compound 5 can be attributed to the acetylation in the hydroquinone moiety.
The predicted values for several pharmacokinetic parameters of compounds 49 related to oral absorption, Caco-2 cell permeability, blood–brain barrier permeability, and binding to human serum albumin, among others, are summarized in Table S3. These ADME descriptor values indicate that the percentage of predicted oral absorption for these compounds ranges from 94% to 100%, suggesting good oral bioavailability. Furthermore, all the evaluated compounds demonstrate good to excellent predicted values for Caco-2 cell permeability, with QPlogBB values falling between −2.6 and −0.68. Additionally, all the tested compounds are within the range of interaction with human serum albumin, suggesting their potential transport by plasma proteins to the target site.
However, it is worth noting that all compounds may block HERG K+-channels, which play a crucial role in cardiac repolarization, potentially increasing the risk of cardiac arrhythmias. Moreover, compound 5 is expected to have sufficient solubility in water, while compounds 49 are considered higher-lipophilicity compounds, enhancing their ability to penetrate cell membranes. In terms of compliance with Jorgensen’s rule of three, most compounds remain within permissible limits, except for PIBHQs 7, 9, and 10, which exhibit three violations.
Moreover, nearly all the evaluated compounds meet Lipinski’s rule of five and its Weber extension criteria, except for compounds 4 and 79 (QPlog/Po/w > 5). However, even these compounds have violations that fall within acceptable limits (Table S4). These results collectively suggest that, from a pharmacokinetic perspective, most of these compounds hold promise as potential candidates for preclinical assays.

3. Materials and Methods

3.1. Chemistry

All chemical reactions were carried out from the starting substrate, compound 1 [65], following the synthesis route to obtain chalcones from precursor 2, as previously described [41]. The reaction progress was monitored by thin-layer chromatography with 60 F254 silica gel plates (Merck, TLC silica gel 60 F254). Column chromatography (CC) purification was performed with silica gel 60 (230–400 mesh, Merck). Melting points were determined on a Stuart SMP 10 apparatus (Stone, Staffordshire, UK) and were uncorrected. IR spectra were recorded on a Perkin Elmer FT IR 1600 spectrophotometer (Perkin Elmer, Norwalk, CN, USA) as a film generated by applying a solution of the derivate in CH2Cl2 on NaCl disks, then by evaporation of the solvent under an IR lamp. 1H NMR spectra were recorded operating at 400.13 MHz, while 13C NMR spectra were recorded at 100.62 MHz on a Bruker Avance 400 Digital NMR spectrometer (Bruker/Analytic, Karlsruhe, Germany) in CDCl3 using TMS as an internal reference. Chemical shift (δ) values are specified as parts per million (ppm), followed by multiplicity and coupling constant (J) in Hertz (Hz). Elemental analyses of C, H, and N were determined by a Perkin Elmer 2400 Series II CHN Elemental Analyzer (Perkin Elmer Inc., Waltham, MA02451, 21,929 USA).

3.1.1. Procedure for the Synthesis and Characterization of Compound 3

Synthesis of (E)-3-(Furan-2-yl)-1-{1-hydroxy-6-(4-methylpent-3-en-1-yl)-4-[(tetrahydro-2H-pyran-2-yl)oxy]-5,8-dihydronaphthalen-2-yl}prop-2-en-1-one (3)

A solution containing precursor 2 (1.00 mmol) and barium hydroxide octahydrate (1.00 mmol) in ethanol (8 mL) was stirred for 10 min. Then, 2-furaldehyde (1.15 mmol) was added, and the mixture was stirred for 30 min at 90 °C. After the reaction, the mixture was added to an ice/water bath, vacuum-filtered, and purified by recrystallization using methanol as the solvent, obtaining pure product 3 as an orange solid (283 mg, 63%): m.p.116–117 °C. IR υmax cm−1 (film) 3374 (O-H), 1639 (C=O), 1583 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.69 (m, 10H, 2CH3, 2CH2, H13, H14, H3′, H4′), 1.94 (m, 2H, CH2, H2′), 2.19 (m, 4H, 2CH2, H10, H9), 3.34 (m, 4H, 2CH2, H8, H5), 3.70 (m, 1H, CH2, H5′), 3.95 (m, 1H, CH2, H5′), 5.19 (t, 1H, J = 6.8 Hz, CH, H11), 5.43 (t, 1H, J = 3.3 Hz, CH, H1′), 5.67 (s, 1H, CH, H7), 6.54 (dd, 1H, J = 3.4 Hz, J = 1.8 Hz, CH, H20), 6.75 (d, 1H, J = 3.4 Hz, CH, H19), 7.50 (d, 2H, J = 15.2 Hz, 2CH, H3, H16), 7.57 (d, 1H, J = 1.5 Hz, CH, H21), 7.68 (d, 1H, J = 15.2 Hz, CH, H17), 13.18 (s, 1H, OH, H1). 13C NMR (CDCl3, TMS, ppm) δ 17.7 (C13), 19.0 (C3′), 24.8 (C8), 25.3 (C4′), 25.7 (C14), 26.2 (C10), 28.4 (C5), 30.7 (C2′), 37.2 (C9), 62.0 (C5′), 97.1 (C1′), 110.7 (C3), 112.8 (C20), 116.3 (C2), 116.6 (C19), 117.7 (C11), 118.2 (C16), 124.1 (C7), 125.0 (C8a), 130.6 (C17), 131.8 (C4a), 133.5 (C12), 135.3 (C6), 145.2 (C21), 146.3 (C4), 151.7 (C18), 156.9 (C1), 192.8 (C15). Elemental analysis calculated for C28H32O5: C, 74.97; H, 7.19. Found: C, 74.99; H, 7.16.

3.1.2. Procedure for the Synthesis and Characterization of Compound 4

Synthesis of (E)-1-[1,4-Dihydroxy-6-(4-methylpent-3-en-1-yl)-5,8-dihydronaphthalen-2-yl]-3-(furan-2-yl) prop-2-en-1-one (4)

Acid monohydrate 4-toluenesulfonic (3.00 mmol) was added to a solution of compound 3 (1.00 mmol) in methanol (12 mL). The reaction mixture was stirred for 3 h at room temperature. After the reaction, the mixture was added to an ice/water bath, vacuum-filtered, and purified by recrystallization using methanol as the solvent, obtaining a red solid (335 mg, 92%): m.p.124–125 °C. IR υmax cm−1 (film) 3381 (O-H), 1634 (C=O), 1581 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.68 (m, 6H, 2CH3, H13, H14), 2.17 (m, 4H, 2CH2, H10, H9), 3.32 (m, 4H, 2CH2, H8, H5), 4.80 (s, 1H, OH, H4), 5.17 (t, 1H, J = 6.7 Hz, CH, H11), 5.68 (s, 1H, CH, H7), 6.54 (s, 1H, CH, H20), 6.74 (d, 1H, J = 3.5 Hz, CH, H19), 7.18 (s, 1H, CH, H3), 7.45 (d, 1H, J = 15.1 Hz, CH, H16), 7.54 (s, 1H, CH, H21), 7.65 (d, 1H, J = 15.1 Hz, CH, H17), 13.00 (s, 1H, OH, H1). 13C NMR (CDCl3, TMS, ppm) δ 17.8 (C13), 24.8 (C8), 25.7 (C14), 26.1 (C10), 28.2 (C5), 37.2 (C9), 110.5 (C3), 112.8 (C20), 116.4 (C2), 116.8 (C19), 117.9 (C11, C16), 124.0 (C7), 125.2 (C8a), 130.6 (C17), 131.9 (C12), 132.7 (C4a), 132.9 (C6), 144.8 (C4), 145.3 (C21), 151.6 (C18), 156.0 (C1), 192.5 (C15). Elemental analysis calculated for C23H24O4: C, 75.80; H, 6.64. Found: C, 75.82; H, 6.65.

3.1.3. Procedure for the Synthesis and Characterization of Compound 5

Synthesis of (E)-2-[3-(Furan-2-yl) prop-2-enoyl]-6-(4-methylpent-3-en-1-yl)-5,8-dihydronaphthalene-1,4-diyl diethanoate (5)

Acetic anhydride (1.00 mL, 10.6 mmol) was added to a solution of compound 4 (0.50 mmol) in pyridine (1.25 mL), and the reaction mixture was maintained in the dark with occasional stirring for 16 h at room temperature. After the reaction, the mixture was added to an ice/water bath, dissolved with CH2Cl2 (30 mL), and subjected to successive extractions with 10% HCl solution (2 × 15 mL) and H2O (2 × 10 mL) until neutral pH of the aqueous phase was achieved. The organic phase was dried with anhydrous Na2SO4, filtered, and evaporated under reduced pressure. The crude product was purified by column chromatography with hexane/ethyl acetate 2:1 as eluent, obtaining an orange solid (215 mg, 96%): m.p.120–121 °C. IR υmax cm−1 (film) 1766 (C=O), 1604 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.67 (m, 6H, 2CH3, H13, H14), 2.26 (m, 10H, 2CH3, 2CH2, H’2, H’4, H10, H9), 3.22 (m, 4H, 2CH2, H5, H8), 5.14 (s, 1H, CH, H11), 5.59 (s, 1H, CH, H7), 6.51 (s, 1H, CH, H20), 6.71 (d, 1H, J = 3.6 Hz, CH, H19), 7.10 (d, 1H, J = 15.5 Hz, CH, H16), 7.35 (s, 1H, CH, H3), 7.43 (d, 1H, J = 15.5 Hz, CH, H17), 7.53 (s, 1H, CH, H21). 13C NMR (CDCl3, TMS, ppm) δ 17.8 (C13), 20.8 (C2′), 20.9 (C4′), 25.3 (C8), 25.7 (C14), 26.0 (C10), 27.9 (C5), 36.9 (C9), 112.7 (C20), 116.6 (C16), 116.9 (C19), 120.7 (C3), 122.1 (C11), 123.8 (C7), 129.9 (C4a), 130.5 (C2), 131.4 (C17), 132.0 (C12), 133.1 (C8a), 133.3 (C6), 144.5 (C4), 145.2 (C21), 145.7 (C1), 151.3 (C18), 169.0 (C3′), 169.1 (C1′), 189.5 (C15). Elemental analysis calculated for C27H28O6: C, 72.30; H, 6.29. Found: C, 72.32; H, 6.27.

3.1.4. General Procedure for the Synthesis and Characterization of Compounds 6, 7, 8, 9, and 10

Hydrazine monohydrate (1.50 mmol) was added to a solution of the respective compounds (4, VII, VIII, IX, and X) (0.50 mmol) in ethanol (10 mL). The reaction mixture was refluxed for 5 h. After the reaction time, the mixture was cooled in an ice/water bath. The resulting solid was extracted with ethyl acetate (2 × 20 mL). After separating the phases, the organic phase was dried with anhydrous Na2SO4, filtered, and evaporated under reduced pressure, yielding the respective impure products of 6, 7, 8, 9, and 10.

Synthesis of 2-[5-(Furan-2-yl)-4,5-dihydro-1H-pyrazol-3-yl]-6-(4-methylpent-3-en-1-yl)-5,8-dihydronaphthalene-1,4-diol (6)

This compound was synthesized by a general procedure using precursor 4 and was purified by recrystallization using ethanol as a solvent, obtaining a yellow solid (136 mg, 71%): m.p.115–116 °C. IR υmax cm−1 (film) 3328 (O-H), 3328 (N-H), 1634 (C=N), 1599 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.67 (m, 6H, 2CH3, H13, H14), 2.16 (m, 4H, 2CH2, H9, H10), 3.29 (m, 6H, 3CH2, H8, H5, H16), 4.88 (t, 1H, J = 8.9 Hz, CH, H17), 5.17 (s, 1H, CH, H11), 5.67 (s, 1H, CH, H7), 6.23 (d, 1H, J = 3.3 Hz, CH, H19), 6.32 (s, 1H, CH, H20), 6.52 (s, 1H, CH, H3), 7.37 (s, 1H, CH, H21), 10.77 (s, 1H, OH, H1). 13C NMR (CDCl3, TMS, ppm) δ 17.8 (C13), 25.2 (C8), 25.7 (C14), 26.2 (C10), 27.6 (C5), 37.3 (C9), 37.9 (C16), 55.5 (C17), 106.4 (C19), 110.3 (C20), 110.4 (C3), 112.5 (C2), 118.1 (C11), 123.5 (C8a), 124.1 (C7), 125.1 (C4a), 131.8 (C12), 133.2 (C6), 142.5 (C21), 145.3 (C4), 149.2 (C1), 154.0 (C18), 155.1 (C15). Elemental analysis calculated for C23H26N2O3: C, 72.99; H, 6.92; N, 7.40. Found: C, 72.97; H, 6.95; N, 7.43.

Synthesis of 6-(4-Methylpent-3-en-1-yl)-2-(5-phenyl-4,5-dihydro-1H-pyrazol-3-yl)-5,8-dihydronaphthalene-1,4-diol (7)

This compound was synthesized by a general procedure using precursor VII and was purified by recrystallization using ethanol as a solvent, obtaining a brown solid (120 mg, 63%): m.p.136–137 °C. IR υmax cm−1 (film) 3341 (O-H), 1627 (C=N), 1600 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.69 (m, 6H, 2CH3, H13, H14), 2.19 (m, 4H, 2CH2, H10, H9), 3.06 (dd, 1H, JAB = 16.4 Hz, JAX = 8.4 Hz, CH2, H16A), 3.37 (m, 5H, 3CH2, H8, H5, H16B), 4.85 (dd, 1H, JBX = 10.7 Hz, JAX = 8.4 Hz, CH, H17X), 5.19 (s, 1H, CH, H11), 5.69 (s, 1H, CH, H7), 6.50 (s, 1H, CH, H3), 7.35 (s, 6H, 5CH, OH, H19, H20, H21, H22, H23, H4), 10.84 (s, 1H, OH, H1). 13C NMR (CDCl3, TMS, ppm) δ 17.2 (C13), 25.2 (C8), 25.7 (C14), 26.2 (C10), 29.7 (C5), 37.3 (C9), 48.6 (C16), 62.6 (C17), 110.2 (C3), 112.2 (C2), 118.2 (C11), 124.1 (C7), 126.3 (C21), 126.9 (C8a), 128.0 (C19, C23), 128.8 (C4a), 129.0 (C20, C22), 131.7 (C12), 133.1 (C6), 142.2 (C18), 145.2 (C4), 149.2 (C1), 154.3 (C15). Elemental analysis calculated for C25H28N2O2: C, 77.29; H, 7.26; N, 7.21. Found: C, 77.33; H, 7.28; N, 7.18.

Synthesis of 2-[5-(4-Methoxyphenyl)-4,5-dihydro-1H-pyrazol-3-yl]-6-(4-methylpent-3-en-1-yl)-5,8-dihydronaphthalene-1,4-diol (8)

This compound was synthesized by a general procedure using precursor VIII and was purified by recrystallization using ethanol as a solvent, obtaining a brown solid (130 mg, 62%): m.p.128–129 °C. IR υmax cm−1 (film) 3376 (O-H), 1667 (C=N), 1612 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.67 (m, 6H, 2CH3, H13, H14), 2.18 (m, 4H, 2CH2, H10, H9), 3.02 (dd, 1H, JAB = 16.5 Hz, JAX = 8.2 Hz, CH2, H16A), 3.21 (m, 2H, CH2, H5), 3.39 (m, 3H, CH2, H8, H16B), 3.81 (s, 3H, CH3, H24), 4.79 (dd, 1H, JBX = 10.2 Hz, JAX = 8.2 Hz, CH, H17X), 5.16 (s, 1H, CH, H11), 5.68 (s, 1H, CH, H7), 6.50 (s, 1H, CH, H3), 6.88 (d, 2H, J = 8.9 Hz, 2CH, H20, H22), 7.26 (d, 3H, J = 8.9 Hz, 2CH, OH, H19, H23, H4), 10.78 (s, 1H, OH, H1). 13C NMR (CDCl3, TMS, ppm) δ 17.7 (C13), 25.2 (C8), 25.7 (C14), 26.2 (C10), 27.6 (C5), 37.3 (C9), 41.5 (C16), 55.3 (C24), 62.2 (C17), 110.2 (C3), 112.8 (C2), 114.3 (C20, C22), 118.2 (C11), 123.4 (C8a), 124.1 (C7), 124.7 (C18), 127.5 (C19, C23), 131.8 (C4a), 133.1 (C12), 134.2 (C6), 145.2 (C4), 149.2 (C1), 154.6 (C15), 159.4 (C21). Elemental analysis calculated for C26H30N2O3: C, 74.61; H, 7.22; N, 6.69. Found: C, 74.65; H, 7.23; N, 6.67.

Synthesis of 6-(4-Methylpent-3-en-1-yl)-2-[5-(4-methylphenyl)-4,5-dihydro-1H-pyrazol-3-yl]-5,8-dihydronaphthalene-1,4-diol (9)

This compound was synthesized by a general procedure using precursor IX and was purified by recrystallization using ethanol as a solvent, obtaining a brown solid (137 mg, 67%): m.p.119–120 °C. IR υmax cm−1 (film) 3340 (O-H), 1634 (C=N), 1600 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.68 (m, 6H, 2CH3, H13, H14), 2.19 (m, 4H, 2CH2, H10, H9), 2.36 (s, 3H, CH3, H24), 3.05 (dd, 1H, JAB = 16.3 Hz, JAX = 8.1 Hz, CH2, H16A), 3.36 (m, 5H, 3CH2, H5, H8, H16B), 4.81 (t, 1H, J = 8.8 Hz, CH, H17X), 5.17 (t, 1H, J = 6.2 Hz, CH, H11), 5.69 (s, 1H, CH, H7), 6.50 (s, 1H, CH, H3), 7.20 (m, 5H, 4CH, OH, H19, H20, H22, H23, H4), 10.85 (s, 1H, OH, H1). 13C NMR (CDCl3, TMS, ppm) δ 17.8 (C13), 21.1 (C24), 25.2 (C8), 25.7 (C14), 26.2 (C10), 27.6 (C5), 37.3 (C9), 41.6 (C16), 62.4 (C17), 110.2 (C3), 112.8 (C2), 118.2 (C11), 123.4 (C8a), 124.1 (C7), 124.7 (C12), 126.2 (C19, C23), 129.6 (C20, C22), 131.8 (C4a), 133.1 (C18), 137.8 (C6), 139.9 (C21), 145.2 (C4), 149.2 (C1), 154.5 (C15). Elemental analysis calculated for C26H30N2O2: C, 77.58; H, 7.51; N, 6.96. Found: C, 77.60; H, 7.47; N, 6.98.

Synthesis of 2-[5-(2,4-Dichlorophenyl)-4,5-dihydro-1H-pyrazol-3-yl]-6-(4-methylpent-3-en-1-yl)-5,8-dihydronaphthalene-1,4-diol (10)

This compound was synthesized by a general procedure using precursor X and was purified by recrystallization using methanol as a solvent, obtaining a brown solid (174 mg, 75%): m.p.138–139 °C. IR υmax cm−1 (film) 3349 (O-H), 1627 (C=N), 1600 (C=C). 1H NMR (CDCl3, TMS, ppm) δ 1.67 (m, 6H, 2CH3, H13, H14), 2.16 (m, 4H, 2CH2, H10, H9), 2.84 (dd, 1H, JAB = 16.7 Hz, JAX = 9.4 Hz, CH2, H16A), 3.27 (m, 4H, 2CH2, H5, H8), 3.61 (dd, 1H, JAB = 16.7 Hz, JBX = 10.7 Hz, CH2, H16B), 5.18 (m, 2H, 2CH, H11, H17X), 5.66 (s, 1H, CH, H7), 6.48 (s, 1H, CH, H3), 7.24 (dd, 1H, J = 8.37 Hz, J = 1.95 Hz, CH, H22), 7.41 (d, 1H, J = 1.94 Hz, CH2, H20), 7.53 (d, 1H, J = 8.77 Hz, CH, H23), 7.66 (s, 1H, OH, H4), 10.76 (s, 1H, OH, H1). 13C NMR (CDCl3, TMS, ppm) δ 17.8 (C13), 25.1 (C8), 25.7 (C14), 26.2 (C10), 27.6 (C5), 37.2 (C9), 40.2 (C16), 58.9 (C17), 110.2 (C3), 112.4 (C2), 118.1 (C11), 123.4 (C8a), 124.1 (C7), 125.1 (C4a), 127.7 (C22), 128.2 (C20), 129.4 (C23), 131.8 (C12), 133.2 (C21), 133.4 (C19), 134.0 (C18), 137.9 (C6), 145.3 (C4), 149.1 (C1), 154.6 (C15). Elemental analysis calculated for C25H26Cl2N2O2: C, 65.65; H, 5.73; N, 6.12. Found: C, 65.68; H, 5.78; N, 6.17.

3.2. Antiproliferative Assay

Human cell lines HT-29 (human colon adenocarcinoma, ATCC HTB-38) and MCF-7 (human breast adenocarcinoma, ATCC HTB-22) were obtained from the American Type Culture Collection (Manassas, VA, USA) and were maintained in Dulbecco’s Modified Eagles Medium (DMEM) and supplemented with 10% FBS, 100 U/mL penicillin, and 100 μg/mL streptomycin. Cells were subcultured, and antiproliferative assays were carried out following the experimental conditions we previously described [38]. The absorbance at 490 nm was recorded using a Varioskan Flash Multimode Reader (Thermo Scientific, Waltham, MA, USA). To obtain IC50 values for each compound, dose–response curves were constructed for both cell lines (MCF-7 and HT-29). Results are reported as the mean value of three independent experiments, with each variation performed in triplicate. Doxorubicin was included as a reference anticancer compound in all evaluations.

3.3. Computational Details

3.3.1. Ligand Preparation

The 3D structure of each compound was prepared using ChemDraw Ultra version 12.0, as previously described [38]. The hydrophobicity index (cLogP), drug-likeness values, and toxicity risks were predicted using DataWarrior algorithms [66,67].

3.3.2. In Silico ADME Prediction

Pharmacokinetic parameters were calculated using QikProp (QP) version 4.3 of the Schrodinger suite, based on Lipinski’s rule of five and its extensions, as previously described [38].

3.3.3. Macromolecule Selection and Retrieval

The crystal structures of 16 selected proteins (Table S1), including growth factor receptors, transcription regulators, and enzymes (such as reductases, oxidases, and kinases) were retrieved from the Protein Data Bank [68]. These proteins are overexpressed in some malignancies, including breast and colon adenocarcinoma, as described in the literature [43,44,45,46,61,69,70,71,72,73,74,75,76,77].

3.3.4. Molecular Docking of Ligand–Protein Interactions

Virtual screening was performed using Autodock Vina, a target-specific scoring method useful for virtual screening [78]. All chalcone–pyrazoline–hydroquinone hybrids were docked into a set of proteins with known 3D structures to identify potential inhibitors. Both ligands and proteins were prepared using AutoDock Tools version 1.5.7 (ADT), as previously described [38,79,80]. The binding sites and energies of each compound were predicted for each receptor using Autodock Vina [78]. Graphic analysis of the molecular coupling studies was performed using Visual Molecular Dynamics 1.9 (VMD) [80] and Discovery Studio Biovia 2021 [81].

4. Conclusions

This study presents a novel series of chalcone-1,4-benzohydroquinones (CBHQs) synthesized by replacing the aryl group of the β-carbon of the chalcone with the furan heterocycle. The synthetic process involved the classic Claisen–Schmidt condensation reaction. Additionally, the α, β-unsaturated carbonyl system of the CBHQs underwent a conjugated addition with hydrazine to yield pyrazoline-1,4-benzohydroquinone hybrids (PIBHQs). CBHQs 4 and 5 displayed superior pIC50 values in MCF-7 and HT-29 cells compared to PIBHQs 69, except for CBHQ 6; however, compound 8 exhibited higher cytotoxic activity against HT-29 cells. This indicates that the chalcone system and the acetylation of aromatic rings enhance the antiproliferative activity of the newly synthesized hybrid derivatives against both the MCF-7 and HT-29 cancer cell lines. Moreover, the substitution of the benzene ring by the furan heterocycle in the chemical skeletons of CBHQ hybrids was crucial for obtaining antineoplastic compounds, as their respective analogues, VIIX, showed inactivity against both cell lines.
From a theoretical standpoint, the binding energy of cancer-related proteins with CBHQs and PIBHQs was found to be generally higher for kinases such as EGFR, MEK1, and c-MET, with ΔGbin values ranging from −11.4 to −8.5 kcal/mol. In this regard, the synthesized hybrids could be promising chemotherapeutic agents, potentially targeting kinase pathways. Furthermore, they could serve as scaffolds for the development of novel multi-target anticancer agents. The favorable predictions of physicochemical and pharmacokinetic parameters for most of these compounds, aligning well with previous in vitro antiproliferative results, underscore their potential as promising candidates for preclinical assays.

Supplementary Materials

The following supporting information can be downloaded at: https://mdpi.longhoe.net/article/10.3390/ijms25137281/s1.

Author Contributions

A.M., A.O., and J.M. conceived, designed, and supervised all the experiments; W.A. and J.M. analyzed the in silico data and performed the theoretical calculations; L.G. performed statistical analysis; J.M. and L.G. supervised the bioassays and analyzed the in vitro data; A.M. and A.O. contributed ideas and analyzed the data; J.M., W.A., and A.M. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

We are grateful for financial support from the Dirección de Investigación de la Vicerrectoría de Investigación y Estudios Avanzados, Pontificia Universidad Católica de Valparaíso, Chile (Projects DI 039.471/2020, DI 039.338/2022 and DI 125.750/2023). This research was partially supported by HPC OCÉANO (FONDEQUIP Nº EQM170214) and the supercomputing infrastructure of the NLHPC (CCSS210001). J. Maldonado is grateful for the support of Doctoral Fellowship CONICYT- PCHA/Doctorado Nacional/2014-21140688.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rebucci, M.; Michiels, C. Molecular Aspects of Cancer Cell Resistance to Chemotherapy. Biochem. Pharmacol. 2013, 85, 1219–1226. [Google Scholar] [CrossRef]
  2. Patel, T.K.; Adhikari, N.; Amin, S.A.; Biswas, S.; Jha, T.; Ghosh, B. Small Molecule Drug Conjugates (SMDCs): An Emerging Strategy for Anticancer Drug Design and Discovery. New J. Chem. 2021, 45, 5291–5321. [Google Scholar] [CrossRef]
  3. Bertanha, C.; Januário, A.; Alvarenga, T.; Pimenta, L.; Silva, M.; Cunha, W.; Pauletti, P. Quinone and Hydroquinone Metabolites from the Ascidians of the Genus Aplidium. Mar. Drugs 2014, 12, 3608–3633. [Google Scholar] [CrossRef]
  4. Giner, R.M.; Ríos, J.L.; Máñez, S. Antioxidant Activity of Natural Hydroquinones. Antioxidants 2022, 11, 343. [Google Scholar] [CrossRef]
  5. Lane, A.L.; Mular, L.; Drenkard, E.J.; Shearer, T.L.; Engel, S.; Fredericq, S.; Fairchild, C.R.; Prudhomme, J.; Le Roch, K.; Hay, M.E.; et al. Ecological Leads for Natural Product Discovery: Novel Sesquiterpene Hydroquinones from the Red Macroalga Peyssonnelia sp. Tetrahedron 2010, 66, 455–461. [Google Scholar] [CrossRef]
  6. Scozzafava, A.; Kalın, P.; Supuran, C.T.; Gülçin, İ.; Alwasel, S.H. The Impact of Hydroquinone on Acetylcholine Esterase and Certain Human Carbonic Anhydrase Isoenzymes (HCA I, II, IX, and XII). J. Enzym. Inhib. Med. Chem. 2015, 30, 941–946. [Google Scholar] [CrossRef]
  7. Tasseau, O.; Mosset, P.; Barillé-Nion, S.; Gautier, F.; Juin, P.; Levoin, N.; Amireddy, N.; Kalivendi, S.; Grée, R. Synthesis and Cytotoxicity Studies of Newly Designed Benzyl-Hydroquinone Derivatives. Med. Chem. Res. 2018, 27, 1050–1065. [Google Scholar] [CrossRef]
  8. Sladic, D.; Gasic, M. Reactivity and Biological Activity of the Marine Sesquiterpene Hydroquinone Avarol and Related Compounds from Sponges of the Order Dictyoceratida. Molecules 2006, 11, 1–33. [Google Scholar] [CrossRef]
  9. Peduto, A.; Scuotto, M.; Krauth, V.; Roviezzo, F.; Rossi, A.; Temml, V.; Esposito, V.; Stuppner, H.; Schuster, D.; D’Agostino, B.; et al. Optimization of Benzoquinone and Hydroquinone Derivatives as Potent Inhibitors of Human 5-Lipoxygenase. Eur. J. Med. Chem. 2017, 127, 715–726. [Google Scholar] [CrossRef]
  10. Chang, H.-S.; Lin, Y.-J.; Lee, S.-J.; Yang, C.-W.; Lin, W.-Y.; Tsai, I.-L.; Chen, I.-S. Cytotoxic Alkyl Benzoquinones and Alkyl Phenols from Ardisia Virens. Phytochemistry 2009, 70, 2064–2071. [Google Scholar] [CrossRef]
  11. Byeon, S.; Yi, Y.-S.; Lee, J.; Yang, W.; Kim, J.; Kim, J.; Hong, S.; Kim, J.-H.; Cho, J. Hydroquinone Exhibits In Vitro and In Vivo Anti-Cancer Activity in Cancer Cells and Mice. Int. J. Mol. Sci. 2018, 19, 903. [Google Scholar] [CrossRef]
  12. Sunassee, S.N.; Davies-Coleman, M.T. Cytotoxic and Antioxidant Marine Prenylated Quinones and Hydroquinones. Nat. Prod. Rep. 2012, 29, 513. [Google Scholar] [CrossRef]
  13. Müller, W.E.; Sladić, D.; Zahn, R.K.; Bässler, K.H.; Dogović, N.; Gerner, H.; Gasić, M.J.; Schröder, H.C. Avarol-Induced DNA Strand Breakage in Vitro and in Friend Erythroleukemia Cells. Cancer Res. 1987, 47, 6565–6571. [Google Scholar]
  14. Matulja, D.; Vranješević, F.; Kolympadi Markovic, M.; Pavelić, S.K.; Marković, D. Anticancer Activities of Marine-Derived Phenolic Compounds and Their Derivatives. Molecules 2022, 27, 1449. [Google Scholar] [CrossRef]
  15. Yoo, H.-D.; Leung, D.; Sanghara, J.; Daley, D.; Soest, R.; Andersen, R.J. Isoarenarol, A New Protein Kinase Inhibitor from the Marine Sponge Dysidea Arenaria. Pharm. Biol. 2003, 41, 223–225. [Google Scholar] [CrossRef]
  16. Nursamsiar; Asnawi, A.; Kartasasmita, R.E.; Ibrahim, S.; Tjahjono, D.H. Synthesis, Biological Evaluation, and Docking Analysis of Methyl Hydroquinone and Bromo Methyl Hydroquinone as Potent Cyclooxygenase (COX-1 and COX-2) Inhibitors. J. Appl. Pharm. Sci. 2018, 8, 16–20. [Google Scholar] [CrossRef]
  17. Rammohan, A.; Reddy, J.S.; Sravya, G.; Rao, C.N.; Zyryanov, G.V. Chalcone Synthesis, Properties and Medicinal Applications: A Review. Environ. Chem. Lett. 2020, 18, 433–458. [Google Scholar] [CrossRef]
  18. Al Zahrani, N.A.; El-Shishtawy, R.M.; Elaasser, M.M.; Asiri, A.M. Synthesis of Novel Chalcone-Based Phenothiazine Derivatives as Antioxidant and Anticancer Agents. Molecules 2020, 25, 4566. [Google Scholar] [CrossRef]
  19. Pereira, R.; Silva, A.M.S.; Ribeiro, D.; Silva, V.L.M.; Fernandes, E. Bis-Chalcones: A Review of Synthetic Methodologies and Anti-Inflammatory Effects. Eur. J. Med. Chem. 2023, 252, 115280. [Google Scholar] [CrossRef]
  20. Rajendran, G.; Bhanu, D.; Aruchamy, B.; Ramani, P.; Pandurangan, N.; Bobba, K.N.; Oh, E.J.; Chung, H.Y.; Gangadaran, P.; Ahn, B.-C. Chalcone: A Promising Bioactive Scaffold in Medicinal Chemistry. Pharmaceuticals 2022, 15, 1250. [Google Scholar] [CrossRef]
  21. Liu, W.; He, M.; Li, Y.; Peng, Z.; Wang, G. A Review on Synthetic Chalcone Derivatives as Tubulin Polymerisation Inhibitors. J. Enzym. Inhib. Med. Chem. 2022, 37, 9–38. [Google Scholar] [CrossRef]
  22. Ouyang, Y.; Li, J.; Chen, X.; Fu, X.; Sun, S.; Wu, Q. Chalcone Derivatives: Role in Anticancer Therapy. Biomolecules 2021, 11, 894. [Google Scholar] [CrossRef]
  23. WalyEldeen, A.A.; Sabet, S.; El-Shorbagy, H.M.; Abdelhamid, I.A.; Ibrahim, S.A. Chalcones: Promising Therapeutic Agents Targeting Key Players and Signaling Pathways Regulating the Hallmarks of Cancer. Chem. Biol. Interact. 2023, 369, 110297. [Google Scholar] [CrossRef]
  24. Constantinescu, T.; Lungu, C.N. Anticancer Activity of Natural and Synthetic Chalcones. Int. J. Mol. Sci. 2021, 22, 11306. [Google Scholar] [CrossRef]
  25. Mezgebe, K.; Melaku, Y.; Mulugeta, E. Synthesis and Pharmacological Activities of Chalcone and Its Derivatives Bearing N -Heterocyclic Scaffolds: A Review. ACS Omega 2023, 8, 19194–19211. [Google Scholar] [CrossRef]
  26. Mathew, B.; Suresh, J.; Anbazghagan, S.; Paulraj, J.; Krishnan, G.K. Heteroaryl Chalcones: Mini Review about Their Therapeutic Voyage. Biomed. Prev. Nutr. 2014, 4, 451–458. [Google Scholar] [CrossRef]
  27. Farhat, J.; Alzyoud, L.; Alwahsh, M.; Al-Omari, B. Structure–Activity Relationship of Benzofuran Derivatives with Potential Anticancer Activity. Cancers 2022, 14, 2196. [Google Scholar] [CrossRef]
  28. Gao, F.; Huang, G.; **ao, 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]
  29. Ardiansah, B. Chalcones Bearing N, O, and S-Heterocycles: Recent Notes on Their Biological Significances. J. Appl. Pharm. Sci. 2019, 9, 117–129. [Google Scholar] [CrossRef]
  30. Haider, K.; Shafeeque, M.; Yahya, S.; Yar, M.S. A Comprehensive Review on Pyrazoline Based Heterocyclic Hybrids as Potent Anticancer Agents. Eur. J. Med. Chem. Rep. 2022, 5, 100042. [Google Scholar] [CrossRef]
  31. Alkamaly, O.M.; Altwaijry, N.; Sabour, R.; Harras, M.F. Dual EGFR/VEGFR2 Inhibitors and Apoptosis Inducers: Synthesis and Antitumor Activity of Novel Pyrazoline Derivatives. Arch. Pharm. 2021, 354, 2000351. [Google Scholar] [CrossRef]
  32. Ahmed, M.F.; Santali, E.Y.; El-Haggar, R. Novel Piperazine–Chalcone Hybrids and Related Pyrazoline Analogues Targeting VEGFR-2 Kinase; Design, Synthesis, Molecular Docking Studies, and Anticancer Evaluation. J. Enzym. Inhib. Med. Chem. 2021, 36, 308–319. [Google Scholar] [CrossRef] [PubMed]
  33. Dofe, V.S.; Sarkate, A.P.; Tiwari, S.V.; Lokwani, D.K.; Karnik, K.S.; Kale, I.A.; Dodamani, S.; Jalalpure, S.S.; Burra, P.V.L.S. Ultrasound Assisted Synthesis of Tetrazole Based Pyrazolines and Isoxazolines as Potent Anticancer Agents via Inhibition of Tubulin Polymerization. Bioorg. Med. Chem. Lett. 2020, 30, 127592. [Google Scholar] [CrossRef]
  34. Song, Y.; Feng, S.; Feng, J.; Dong, J.; Yang, K.; Liu, Z.; Qiao, X. Synthesis and Biological Evaluation of Novel Pyrazoline Derivatives Containing Indole Skeleton as Anti-Cancer Agents Targeting Topoisomerase II. Eur. J. Med. Chem. 2020, 200, 112459. [Google Scholar] [CrossRef]
  35. Singh, A.K.; Kumar, A.; Singh, H.; Sonawane, P.; Paliwal, H.; Thareja, S.; Pathak, P.; Grishina, M.; Jaremko, M.; Emwas, A.-H.; et al. Concept of Hybrid Drugs and Recent Advancements in Anticancer Hybrids. Pharmaceuticals 2022, 15, 1071. [Google Scholar] [CrossRef]
  36. Mancini, I.; Vigna, J.; Sighel, D.; Defant, A. Hybrid Molecules Containing Naphthoquinone and Quinolinedione Scaffolds as Antineoplastic Agents. Molecules 2022, 27, 4948. [Google Scholar] [CrossRef]
  37. Molinari, A.; Oliva, A.; Arismendi, M.; Imbarack, E.; Gálvez, C.; Maldonado, J.; Feliciano, A.S. The Synthesis of Some Fused Pyrazolo-1,4-Naphthoquinones. J. Heterocycl. Chem. 2015, 52, 620–622. [Google Scholar] [CrossRef]
  38. Maldonado, J.; Acevedo, W.; Molinari, A.; Oliva, A.; Knox, M.; San Feliciano, A. Synthesis, in Vitro Evaluation and Molecular Docking Studies of Novel Naphthoisoxazolequinone Carboxamide Hybrids as Potential Antitumor Agents. Polycycl. Aromat. Compd. 2023, 43, 4960–4983. [Google Scholar] [CrossRef]
  39. Molinari, A.; Oliva, A.; Arismendi-Macuer, M.; Guzmán, L.; Fuentealba, M.; Knox, M.; Vinet, R.; San Feliciano, A. New 1H-Benzo[f]Indazole-4,9-Diones Conjugated with C-Protected Amino Acids and Other Derivatives: Synthesis and in Vitro Antiproliferative Evaluation. Molecules 2015, 20, 21924–21938. [Google Scholar] [CrossRef]
  40. Molinari, A.; Oliva, A.; Arismendi-Macuer, M.; Guzmán, L.; Acevedo, W.; Aguayo, D.; Vinet, R.; San Feliciano, A. Antiproliferative Benzoindazolequinones as Potential Cyclooxygenase-2 Inhibitors. Molecules 2019, 24, 2261. [Google Scholar] [CrossRef]
  41. Maldonado, J.; Oliva, A.; Molinari, A.; Acevedo, W. 2-Acetyl-5,8-Dihydro-6-(4-Methyl-3-Pentenyl)-1,4-Naphthohydroquinone-Derived Chalcones as Potential Anticancer Agents. Molecules 2023, 28, 7172. [Google Scholar] [CrossRef]
  42. Guo, S.; Colbert, L.S.; Fuller, M.; Zhang, Y.; Gonzalez-Perez, R.R. Vascular Endothelial Growth Factor Receptor-2 in Breast Cancer. Biochim. Biophys. Acta-Rev. Cancer 2010, 1806, 108–121. [Google Scholar] [CrossRef]
  43. Ali, S.; Coombes, R.C. Estrogen Receptor Alpha in Human Breast Cancer: Occurrence and Significance. J. Mammary Gland Biol. Neoplasia 2000, 5, 271–281. [Google Scholar] [CrossRef]
  44. Kim, J.; Kim, S.; Ko, S.; In, Y.; Moon, H.; Ahn, S.K.; Kim, M.K.; Lee, M.; Hwang, J.; Ju, Y.S.; et al. Recurrent Fusion Transcripts Detected by Whole-transcriptome Sequencing of 120 Primary Breast Cancer Samples. Genes Chromosom. Cancer 2015, 54, 681–691. [Google Scholar] [CrossRef]
  45. Ahmad, D.A.J.; Negm, O.H.; Alabdullah, M.L.; Mirza, S.; Hamed, M.R.; Band, V.; Green, A.R.; Ellis, I.O.; Rakha, E.A. Clinicopathological and Prognostic Significance of Mitogen-Activated Protein Kinases (MAPK) in Breast Cancers. Breast Cancer Res. Treat. 2016, 159, 457–467. [Google Scholar] [CrossRef]
  46. Pashirzad, M.; Khorasanian, R.; Fard, M.M.; Arjmand, M.-H.; Langari, H.; Khazaei, M.; Soleimanpour, S.; Rezayi, M.; Ferns, G.A.; Hassanian, S.M.; et al. The Therapeutic Potential of MAPK/ERK Inhibitors in the Treatment of Colorectal Cancer. Curr. Cancer Drug Targets 2021, 21, 932–943. [Google Scholar] [CrossRef]
  47. Sharma, B.; Singh, V.J.; Chawla, P.A. Epidermal Growth Factor Receptor Inhibitors as Potential Anticancer Agents: An Update of Recent Progress. Bioorg. Chem. 2021, 116, 105393. [Google Scholar] [CrossRef]
  48. Nawaz, F.; Alam, O.; Perwez, A.; Rizvi, M.A.; Naim, M.J.; Siddiqui, N.; Ul Firdaus, J.; Rahman, S.; Jha, M.; Sheikh, A.A. Design, Synthesis, Molecular Docking, and Anticancer Evaluation of Pyrazole Linked Pyrazoline Derivatives with Carbothioamide Tail as EGFR Kinase Inhibitors. Anticancer Agents Med. Chem. 2020, 21, 42–60. [Google Scholar] [CrossRef]
  49. Raghu, M.S.; Pradeep Kumar, C.B.; Prashanth, M.K.; Yogesh Kumar, K.; Prathibha, B.S.; Kanthimathi, G.; Alissa, S.A.; Alghulikah, H.A.; Osman, S.M. Novel 1,3,5-Triazine-Based Pyrazole Derivatives as Potential Antitumor Agents and EFGR Kinase Inhibitors: Synthesis, Cytotoxicity, DNA Binding, Molecular Docking and DFT Studies. New J. Chem. 2021, 45, 13909–13924. [Google Scholar] [CrossRef]
  50. Al-Anazi, M.; Khairuddean, M.; Al-Najjar, B.O.; Murwih Alidmat, M.; Nur Syazni Nik Mohamed Kamal, N.; Muhamad, M. Synthesis, Anticancer Activity and Docking Studies of Pyrazoline and Pyrimidine Derivatives as Potential Epidermal Growth Factor Receptor (EGFR) Inhibitors. Arab. J. Chem. 2022, 15, 103864. [Google Scholar] [CrossRef]
  51. Abourehab, M.A.S.; Alqahtani, A.M.; Youssif, B.G.M.; Gouda, A.M. Globally Approved EGFR Inhibitors: Insights into Their Syntheses, Target Kinases, Biological Activities, Receptor Interactions, and Metabolism. Molecules 2021, 26, 6677. [Google Scholar] [CrossRef] [PubMed]
  52. Ge, X.; Zhang, Y.; Huang, F.; Wu, Y.; Pang, J.; Li, X.; Fan, F.; Liu, H.; Li, S. EGFR Tyrosine Kinase Inhibitor Almonertinib Induces Apoptosis and Autophagy Mediated by Reactive Oxygen Species in Non-Small Cell Lung Cancer Cells. Hum. Exp. Toxicol. 2021, 40, S49–S62. [Google Scholar] [CrossRef] [PubMed]
  53. Shen, G.; Zheng, F.; Ren, D.; Du, F.; Dong, Q.; Wang, Z.; Zhao, F.; Ahmad, R.; Zhao, J. Anlotinib: A Novel Multi-Targeting Tyrosine Kinase Inhibitor in Clinical Development. J. Hematol. Oncol. 2018, 11, 120. [Google Scholar] [CrossRef] [PubMed]
  54. Hu, H.; Liu, Y.; Tan, S.; **e, X.X.; He, J.; Luo, F.; Wang, L. Anlotinib Exerts Anti-Cancer Effects on KRAS-Mutated Lung Cancer Cell Through Suppressing the MEK/ERK Pathway. Cancer Manag. Res. 2020, 12, 3579–3587. [Google Scholar] [CrossRef]
  55. Rodig, S.J.; Shapiro, G.I. Crizotinib, a Small-Molecule Dual Inhibitor of the c-Met and ALK Receptor Tyrosine Kinases. Curr. Opin. Investig. Drugs 2010, 11, 1477–1490. [Google Scholar]
  56. Tang, P.A.; Tsao, M.-S.; Moore, M.J. A Review of Erlotinib and Its Clinical Use. Expert Opin. Pharmacother. 2006, 7, 177–193. [Google Scholar] [CrossRef]
  57. Neal, J.W.; Dahlberg, S.E.; Wakelee, H.A.; Aisner, S.C.; Bowden, M.; Huang, Y.; Carbone, D.P.; Gerstner, G.J.; Lerner, R.E.; Rubin, J.L.; et al. Erlotinib, Cabozantinib, or Erlotinib plus Cabozantinib as Second-Line or Third-Line Treatment of Patients with EGFR Wild-Type Advanced Non-Small-Cell Lung Cancer (ECOG-ACRIN 1512): A Randomised, Controlled, Open-Label, Multicentre, Phase 2 Trial. Lancet Oncol. 2016, 17, 1661–1671. [Google Scholar] [CrossRef]
  58. Tsao, M.-S.; Sakurada, A.; Cutz, J.-C.; Zhu, C.-Q.; Kamel-Reid, S.; Squire, J.; Lorimer, I.; Zhang, T.; Liu, N.; Daneshmand, M.; et al. Erlotinib in Lung Cancer—Molecular and Clinical Predictors of Outcome. N. Engl. J. Med. 2005, 353, 133–144. [Google Scholar] [CrossRef]
  59. Bhangoo, M.S.; Sigal, D. TRK Inhibitors: Clinical Development of Larotrectinib. Curr. Oncol. Rep. 2019, 21, 14. [Google Scholar] [CrossRef]
  60. Federman, N.; McDermott, R. Larotrectinib, a Highly Selective Tropomyosin Receptor Kinase (TRK) Inhibitor for the Treatment of TRK Fusion Cancer. Expert Rev. Clin. Pharmacol. 2019, 12, 931–939. [Google Scholar] [CrossRef]
  61. Zhang, X.; Li, B.; Song, M.; Song, J. Expression and Significance of ERK Protein in Human Breast Carcinoma. Chin. J. Cancer Res. 2004, 16, 269–273. [Google Scholar] [CrossRef]
  62. Ho-Yen, C.M.; Jones, J.L.; Kermorgant, S. The Clinical and Functional Significance of C-Met in Breast Cancer: A Review. Breast Cancer Res. 2015, 17, 52. [Google Scholar] [CrossRef] [PubMed]
  63. Griffin, N.; Marsland, M.; Roselli, S.; Oldmeadow, C.; Attia, J.; Walker, M.M.; Hondermarck, H.; Faulkner, S. The Receptor Tyrosine Kinase TrkA Is Increased and Targetable in HER2-Positive Breast Cancer. Biomolecules 2020, 10, 1329. [Google Scholar] [CrossRef] [PubMed]
  64. Brylinski, M. Aromatic Interactions at the Ligand–Protein Interface: Implications for the Development of Docking Scoring Functions. Chem. Biol. Drug Des. 2018, 91, 380–390. [Google Scholar] [CrossRef] [PubMed]
  65. Molinari, A.; Oliva, A.; del Corral, J.M.M.; Castro, M.A.; Araya, C.; García-Grávalos, M.D.; San Feliciano, A. Cytotoxic–Antineoplastic Activity of Acetyl Derivatives of Prenylnaphthohydroquinone. Il Farm. 2004, 59, 651–656. [Google Scholar] [CrossRef] [PubMed]
  66. Sander, T.; Freyss, J.; von Korff, M.; Rufener, C. DataWarrior: An Open-Source Program For Chemistry Aware Data Visualization And Analysis. J. Chem. Inf. Model. 2015, 55, 460–473. [Google Scholar] [CrossRef] [PubMed]
  67. Jorgensen, W.L.; Duffy, E.M. Prediction of Drug Solubility from Structure. Adv. Drug Deliv. Rev. 2002, 54, 355–366. [Google Scholar] [CrossRef] [PubMed]
  68. Berman, H.M.; Battistuz, T.; Bhat, T.N.; Bluhm, W.F.; Bourne, P.E.; Burkhardt, K.; Feng, Z.; Gilliland, G.L.; Iype, L.; Jain, S.; et al. The Protein Data Bank. Acta Crystallogr. Sect. D Biol. Crystallogr. 2002, 58, 899–907. [Google Scholar] [CrossRef] [PubMed]
  69. Negi, R.R.; Rana, S.V.; Gupta, V.; Gupta, R.; Chadha, V.D.; Prasad, K.K.; Dhawan, D.K. Over-Expression of Cyclooxygenase-2 in Colorectal Cancer Patients. Asian Pac. J. Cancer Prev. 2019, 20, 1675–1681. [Google Scholar] [CrossRef]
  70. Sylvester, P.W. Targeting Met Mediated Epithelial-mesenchymal Transition in the Treatment of Breast Cancer. Clin. Transl. Med. 2014, 3, 30. [Google Scholar] [CrossRef]
  71. Wang, Y.; Zhou, B.P. Epithelial-Mesenchymal Transition in Breast Cancer Progression and Metastasis. Chin. J. Cancer 2011, 30, 603–611. [Google Scholar] [CrossRef] [PubMed]
  72. Lazaro, G.; Smith, C.; Goddard, L.; Jordan, N.; McClelland, R.; Barrett-Lee, P.; Nicholson, R.I.; Hiscox, S. Targeting Focal Adhesion Kinase in ER+/HER2+ Breast Cancer Improves Trastuzumab Response. Endocr. Relat. Cancer 2013, 20, 691–704. [Google Scholar] [CrossRef] [PubMed]
  73. Nakano, M.; Fukami, T.; Gotoh, S.; Nakajima, M. A-to-I RNA Editing Up-Regulates Human Dihydrofolate Reductase in Breast Cancer. J. Biol. Chem. 2017, 292, 4873–4884. [Google Scholar] [CrossRef] [PubMed]
  74. Lei, H.; Deng, C.-X. Fibroblast Growth Factor Receptor 2 Signaling in Breast Cancer. Int. J. Biol. Sci. 2017, 13, 1163–1171. [Google Scholar] [CrossRef] [PubMed]
  75. Haldosén, L.-A.; Zhao, C.; Dahlman-Wright, K. Estrogen Receptor Beta in Breast Cancer. Mol. Cell. Endocrinol. 2014, 382, 665–672. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, H.; Zhu, G.; Li, Y.; Padia, R.N.; Dong, Z.; Pan, Z.K.; Liu, K.; Huang, S. Extracellular Signal–Regulated Kinase Signaling Pathway Regulates Breast Cancer Cell Migration by Maintaining Slug Expression. Cancer Res. 2009, 69, 9228–9235. [Google Scholar] [CrossRef] [PubMed]
  77. Sultan, A.S.; Brim, H.; Sherif, Z.A. Co-overexpression of Janus Kinase 2 and Signal Transducer and Activator of Transcription 5a Promotes Differentiation of Mammary Cancer Cells through Reversal of Epithelial–Mesenchymal Transition. Cancer Sci. 2008, 99, 272–279. [Google Scholar] [CrossRef] [PubMed]
  78. Trott, O.; Olson, A.J. AutoDock Vina: Improving the Speed and Accuracy of Docking with a New Scoring Function, Efficient Optimization, and Multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef] [PubMed]
  79. Acevedo, W.; González-Nilo, F.; Agosin, E. Docking and Molecular Dynamics of Steviol Glycoside–Human Bitter Receptor Interactions. J. Agric. Food Chem. 2016, 64, 7585–7596. [Google Scholar] [CrossRef]
  80. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  81. Dassault Syst èmes BIOVIA. Discovery Studio Visualizer; V20.1.0, Vol19295; Dassault Systèmes: San Diego, CA, USA, 2019. [Google Scholar]
Figure 1. Design of novel hydroquinone hybrids as anticancer agents.
Figure 1. Design of novel hydroquinone hybrids as anticancer agents.
Ijms 25 07281 g001
Scheme 1. Synthesis route of hybrid 4. Reagents and conditions: (a) DHP, PPTS, CH2Cl2, rt, 19 h; (b) Ba(OH)2·8H2O, 2-furaldehyde, EtOH, 90 °C, 30 min; (c) PTSA, MeOH, rt, 3 h.
Scheme 1. Synthesis route of hybrid 4. Reagents and conditions: (a) DHP, PPTS, CH2Cl2, rt, 19 h; (b) Ba(OH)2·8H2O, 2-furaldehyde, EtOH, 90 °C, 30 min; (c) PTSA, MeOH, rt, 3 h.
Ijms 25 07281 sch001
Scheme 2. Synthesis of hybrid molecules 5 and 6. Reagents and conditions: (a) acetic anhydride, pyridine, rt, 24 h; (b) N2H4·H2O, EtOH, reflux, 5 h.
Scheme 2. Synthesis of hybrid molecules 5 and 6. Reagents and conditions: (a) acetic anhydride, pyridine, rt, 24 h; (b) N2H4·H2O, EtOH, reflux, 5 h.
Ijms 25 07281 sch002
Scheme 3. Synthesis of hybrid molecules 7, 8, 9, and 10. Reagents and conditions: (a) N2H4·H2O, EtOH, reflux, 5 h.
Scheme 3. Synthesis of hybrid molecules 7, 8, 9, and 10. Reagents and conditions: (a) N2H4·H2O, EtOH, reflux, 5 h.
Ijms 25 07281 sch003
Figure 2. Antiproliferative effects of isosteric replacement and pyrazoline heterocycle incorporation into synthesized chalcones.
Figure 2. Antiproliferative effects of isosteric replacement and pyrazoline heterocycle incorporation into synthesized chalcones.
Ijms 25 07281 g002
Figure 3. Structure–activity relationship (SAR) of chalcone hybrids and their synthesized pyrazoline derivatives.
Figure 3. Structure–activity relationship (SAR) of chalcone hybrids and their synthesized pyrazoline derivatives.
Ijms 25 07281 g003
Figure 4. Two-dimensional maps of H bonds and hydrophobic interactions of CNQ 6, 7, 8, and 9 with EGFR residues. Van der Waals, Pi–anion, Pi–sigma, Pi–Pi stacked, and Pi–alkyl interactions are considered hydrophobic.
Figure 4. Two-dimensional maps of H bonds and hydrophobic interactions of CNQ 6, 7, 8, and 9 with EGFR residues. Van der Waals, Pi–anion, Pi–sigma, Pi–Pi stacked, and Pi–alkyl interactions are considered hydrophobic.
Ijms 25 07281 g004
Figure 5. (A) Visualization of the potential binding site of PBHQ hybrid 7 into EGFR; (B) detail of its H bonding with Ala743, Lys745, and Leu788; (C) superimposition of the docking poses for PHBQ hybrids 6 (yellow), 7 (green), 8 (cyan), and 9 (purple).
Figure 5. (A) Visualization of the potential binding site of PBHQ hybrid 7 into EGFR; (B) detail of its H bonding with Ala743, Lys745, and Leu788; (C) superimposition of the docking poses for PHBQ hybrids 6 (yellow), 7 (green), 8 (cyan), and 9 (purple).
Ijms 25 07281 g005
Figure 6. (A): Overlap** of the docking poses for PIBHQ hybrid 7 (green), ligand 1 (yellow), and erlotinib (grey) into c-MET. Superimposition of the docking poses for (B) 6c and ligand 1, as well as (C) 6c and erlotinib.
Figure 6. (A): Overlap** of the docking poses for PIBHQ hybrid 7 (green), ligand 1 (yellow), and erlotinib (grey) into c-MET. Superimposition of the docking poses for (B) 6c and ligand 1, as well as (C) 6c and erlotinib.
Ijms 25 07281 g006
Table 1. In vitro cytotoxicity data for compounds 410 in MCF-7 breast adenocarcinoma and HT-29 colon adenocarcinoma cells.
Table 1. In vitro cytotoxicity data for compounds 410 in MCF-7 breast adenocarcinoma and HT-29 colon adenocarcinoma cells.
CompoundsMCF-7HT-29
IC50, μM [a]pIC50 [b]IC50, μMpIC50
428.8 ± 0.844.5433.8 ± 0.944.47
521.8 ± 0.414.6618.7 ± 0.304.73
634.3 ± 1.134.4637.6 ± 0.914.42
7124.6 ± 1.313.90113.0 ± 0.923.95
881.2 ± 0.634.0930.6 ± 0.314.51
999.3 ± 0.924.0089.8 ± 0.724.05
10>300->300-
VII [c]>300->300-
VIII [c]>300->300-
IX [c]>300->300-
X [c]>300->300-
Doxorubicin0.27 ± 0.086.574.07 ± 0.925.39
[a] IC50 (half-maximal inhibitory concentration) mean ± SD values; [b] pIC50 = −log IC50(M); bold: significant cytotoxic effect, IC50 < 40 µM; IC50 > 300 µM: inactive compounds; [c]: CBHQ compounds (taken from reference [41]).
Table 2. Comparison (ΔGbin, kcal/mol) of synthesized cytotoxic hybrids with kinase proteins overexpressed in cancer and kinase inhibitors approved by the FDA.
Table 2. Comparison (ΔGbin, kcal/mol) of synthesized cytotoxic hybrids with kinase proteins overexpressed in cancer and kinase inhibitors approved by the FDA.
CompoundTarget Proteins
EGFRMEK1c-MET TPKTRKACK4
4−10.7−9.6−9.7−8.6−9.2−8.4
5−8.5−8.5−9.5−8.5−9.2−8.3
6−11.3−10.0−10.3−10.0−9.5−7.9
7−11.4−10.7−10.3−10.4−10.0−8.7
8−11.0−10.3−9.7−10.2−9.3−8.8
9−11.2−10.6−9.9−10.6−10.1−9.0
P avge.−10.68−9.95−9.90−9.72−9.55−8.52
Erlotinib−8.6−8.0−9.1−8.5−8.8−7.2
Anlotinib−9.2−9.7−9.0−9.0−10.7−8.5
Crizotinib−9.7−9.9−10.0−10.2−10.1−8.0
Proteins with their respective (PDB) entries. EGFR: epidermal growth factor receptor (5GTY); MEK1: MAPK/ERK kinase (4AN3); c-MET: mesenchymal–epithelial transition factor (3RHK); TPK: tyrosine protein kinase (4EHZ); TRKA: tropomyosin receptor kinase A (6PL2); CK4: cyclin-dependent kinase 4 (1PXL). Bold: Average ΔGbin value for each protein.
Table 3. Predicted binding free-energy values (ΔGbin, kcal/mol) and binding site contacts of synthesized cytotoxic hybrids with amino acids of EGFR, MEK1, CK4, and TPK.
Table 3. Predicted binding free-energy values (ΔGbin, kcal/mol) and binding site contacts of synthesized cytotoxic hybrids with amino acids of EGFR, MEK1, CK4, and TPK.
CompoundΔGbin H Bonds and Hydrophobic Contacts in the Binding Site *
EGFR (mean ΔGbin = −10.68 kcal/mol)
4−10.7Leu718, Gly719, Ser720, Val726 */Ala743 *, Ile744, Lys745 *, Met766, Ala767, Val769, Cys775, Arg776, Leu777, Leu788, Ile789, Met790 */Leu792, Met793, Gly796, Cys797 */Arg841, Leu844, Thr854, Asp855, Phe856, Leu858
5−8.5Leu718, Gly719, Ser720, Gly721, Val726 */Ala743 *, Ile744, Lys745 *, Met766 *, Cys775, Arg776, Leu777 *, Leu788 *, Ile789, Met790 */Gln791, Leu792, Met793, Gly796, Cys797/Asp800, Arg841, Asn842, Leu844, Thr854, Asp855, Leu858
6−11.3Leu718 *, Val726 */Ala743 *, Ile744, Lys745 *, Val 769, Cys775, Met766, Arg776, Leu777, Leu788 *, Ile789, Met790 */Leu792, Met793, Gly796, Cys797 */Asp800, Leu844, Thr854, Asp855, Phe856, Leu858
7−11.4Leu718 *, Val726 */Ala743 *, Ile744, Lys745 *, Met766, Val769, Cys775, Arg776, Leu777, Leu788, Ile789, Met790 */Leu792, Met793, Gly796, Cys797 *, Asp800, Leu844, Thr854, Asp855, Phe856, Leu858
8−11.0Leu718, Gly719, Val726 */Ala743, Ile744, Lys745 *, Met766, Cys775, Arg776, Leu777, Leu788 *, Ile789, Met790 *,/Leu792, Met793, Gly796, Cys797 */Asp800, Arg841, Leu844, Ile853, Thr854, Asp855, Phe856 *, Leu858
9−11.2Leu718 *, Gly719, Val726 */Ala743 *, Ile744, Lys745, Met766, Val769, Cys775, Arg776, Leu777, Leu788 *, Ile789, Met790 */Leu792, Met793, Gly796, Cys797 */Asp800, Leu844, Ile853, Thr854, Asp855, Phe856, Leu858
MEK1 (mean ΔGbin = −9.95 kcal/mol)
4−9.6Lys97, Ile99, Leu115, Leu118, Cys121/Ile126, Val127, Gly128, Phe129/Ile141, Met143, Arg189 *, Asp190/Leu206, Cys207, Asp208, Phe209, Gly210, Val211, Ser212, Leu215 *, Ile216 *, Met219
5−8.5Asn78, Lys97, Ile99, Leu115, Leu118 */Ile141, Met143 *, His188, Arg189, Asp190/Cys207, Asp208, Phe209 *, Gly210, Val211, Ser212, Leu215, Ile216, Met219 */Ala220, Gly225, Thr226, Met230, Arg234
6−10.0Lys97, Leu115, Leu118 */Ile126, Val127, Gly128, Phe129/Ile141, Met143 *, Arg189, Asp190/Leu206, Cys207 *, Asp208 *, Phe209 *, Gly210, Val211, Ser212, Leu215 *, Ile216, Met219/Ala220, Gly225, Thr226, Met230, Arg234
7−10.7Lys97, Leu115, Leu118 */Ile126, Val127, Gly128, Phe129/Ile141, Met143 *, Glu144, Arg189, Asp190/Leu206, Cys207 *, Asp208, Phe209, Gly210, Val211, Ser212, Leu215 *, Ile216, Met219/Ala220, Thr226, Met230, Arg234
8−10.3Lys97, Ile99, Leu115, Leu118 */Val127, Gly128, Phe129/Ile141 *, Met143 *, Arg189, Asp190/Cys207, Asp208 *, Phe209 *, Gly210, Val211, Leu215, Ile216, Met219 */Ala220, Met230, Arg234, Tyr240
9−10.6Lys97, Ile99, Leu115, Leu118 */Ile126, Val127, Gly128, Phe129/Ile141 *, Met143 *, Arg189, Asp190/Leu206, Cys207 *, Asp208 *, Phe209 *, Gly210, Val211, Ser212, Leu215 *, Ile216, Met219 */Ala220, Met230, Arg234, Tyr240
c-MET (mean ΔGbin = −9.82 kcal/mol)
4−9.7Ile1084, Gly1085, His1088, Phe1089 *, Val1092/Ala1108, Val1109, Lys1110, Val1155, Leu1157, Gly1163, Asp1164, Asn1167/Arg1208, Met1211, Phe1223, Ala1226, Arg1227 *, Met1229/Tyr1230, Asp1231, Tyr1234, Tyr1235
5−9.5Ile1084, Gly1085, Gly1087, His1088, Phe1089, Val1092/Ala1108, Leu1140, Leu1157, Gly1163, Asp1164 *, Asn1167/Arg1208, Met1211, Phe1223, Arg1227 *, Met1229/Tyr1230, Asp1231, Tyr1234, Tyr1235
6−10.3Ile1084, Gly1085, Phe1089 *, Val1092/Ala1108, Lys1110, Val1155, Leu1157, His1162, Gly1163, Asp1164, Asn1167/Phe1168, Asn1171, His1174, Arg1208, Met1211, Phe1223, Ala1226, Arg1227
7−10.3Gly1085, Phe1089 *, Val1092, Ala1108, Lys1110, Leu1112, Val1155, Leu1157, His1162, Gly1163, Asp1164, Asn1167, Phe1168, Asn1171, His1174, Arg1208, Met1211, Phe1223, Ala1226, Arg1227
8−9.7Ile1084, Phe1089 *, Val1092/Ala1108, Leu1140, Leu1157, Pro1158, Tyr1159, Met1160, Gly1163, Asp1164, Arg1166, Asn1167 */Asn1208 *, Met1211, Phe1223, Arg1227/Tyr1234
9−9.9Gly1087, His1088, Phe1089, Val1092 */Ala1108 *, Lys1110 *, Leu1112, Val1155, Leu1157 *, Asp1164 *, Arg1166/Asp1167, Arg1208, Met1211, Phe1223 *, Ala1226, Arg1227, Met1229/Tyr1230, Asp1231, Tyr1234, Tyr1235
Bolded names correspond to amino acids involved in H bonds with the corresponding synthesized cytotoxic hybrids. Partial interacting peptide sequences are separated and differentiated by colors to facilitate comparisons of similar interactions. Residues with * correspond to amino acids that interact with the ligand through any type of Pi interaction.
Table 4. Binding site contacts of compound 7, ligands, and drugs into EGFR, MEK1, and c-MET.
Table 4. Binding site contacts of compound 7, ligands, and drugs into EGFR, MEK1, and c-MET.
CompoundΔGbin
(kcal/mol)
H Bonds and Hydrophobic Contacts in the Binding Site
EGFR
7−11.4Leu718, Val726, Ala743, Ile744, Lys745, Met766, Val769, Cys775, Arg776, Leu777, Leu788, Ile789, Met790, Leu792, Met793, Gly796, Cys797, Asp800, Leu844, Thr854, Asp855, Phe856, Leu858
Ligand 1 [a]−10.7Leu718, Gly719, Ser720, Val726, Ala743, Ile744, Lys745, Met766, Cys775, Arg776, Leu777, Leu788, Ile789, Met790, Gln791, Leu792, Met793, Gly796, Cys797, Asp800, Arg841, Leu844, Thr854, Asp855, Phe856, Leu858
Erlotinib [b]−8.6Leu718, Gly721, Val726, Ala743, Lys745, Met766, Cys775, Arg776, Leu777, Leu788, Met790, Met793, Gly796, Arg841, Arg842, Leu844, Thr854, Asp855, Phe856, Leu858
MEK1
7−10.7Lys97, Leu115, Leu118, Ile126, Val127, Gly128, Phe129, Ile141, Met143, Glu144, Arg189, Asp190, Leu206, Cys207, Asp208, Phe209, Gly210, Val211, Ser212, Leu215, Ile216, Met219, Ala220, Thr226, Met230, Arg234
Ligand 2 [a]−11.4Asn78, Gly79, Lys97, Ile99, Leu115, Leu118, Val127, Phe129, Ile141, Met143, Asp190, Cys207, Asp208, Phe209, Gly210, Val211, Ser212, Leu215, Ile216, Met219
Anlotinib [b]−9.7Lys97, Leu115, Leu118, Val127, Gly128, Phe129, Ile141, Met143, Arg189, Asp190, Cys207, Asp208, Phe209, Gly210, Val211, Ser212, Leu215, Ile216, Met219, Gly225
c-MET
7−10.3Gly1085, Phe1089, Val1092, Ala1108, Lys1110, Leu1112, Val1155, Leu1157, His1162, Gly1163, Asp1164, Asn1167, Phe1168, Asn1171, His1174, Arg1208, Met1211, Phe1223, Ala1226, Arg1227
Ligand 3 [a]−14.6Ile1084, Gly1085, Phe1089, Val1092, Ala1108, Lys1110, Leu1140, Leu1157, Met1158, Tyr1159, Met1160, Gly1163, Met1211, Phe1223, Ala1226, Arg1227
Crizotinib [b]−10.0Ile1084, Gly1085, Phe1089, Val1092, Ala1108, Lys1110, Arg1227, Leu1157, Asp1164, Arg1166, Asn1167, Arg1208, Met1211, Phe1223, Ala1226, Tyr1234
[a] Ligands 1, 2, and 3 correspond to 1-[(3R)-3-[4-azanyl-3-[3-chloranyl-4-[(6-methylpyridin-2-yl)methoxy]phenyl]pyrazolo [3,4-d]pyrimidin-1-yl]piperidin-1-yl]prop-2-en-1-one, N-[(2S)-2,3-bis(oxidanyl)propoxy]-3,4-bis(fluoranyl)-2-[(2-fluoranyl-4-iodanyl-phenyl)amino]benzamide, and 2-methyl-1-(piperidin-4-yl)-1,6-dihydroimidazo [4,5-d]pyrrolo [2,3-b]pyridine, respectively. Three-dimensional structures of ligands 1, 2, and 3 were extracted from the Protein Data Bank using PDB IDs 5GTY, 4AN3, and 3RHK, respectively. [b] Drug compounds that act as inhibitors of biological targets. Words colored in blue, green, and red correspond to amino acids shared by 7 and ligands; 7 and drugs; and 7, ligands, and drugs, respectively. Bolded names correspond to amino acids involved in H bonds and 7–enzyme interactions.
Table 5. Comparison of predicted toxicity risks a and drug-likeness scores a for hybrids 49.
Table 5. Comparison of predicted toxicity risks a and drug-likeness scores a for hybrids 49.
CompoundMTIRDrug Likeness
4hnnn−1.58
5hnhn−1.49
6nnnn−0.47
7nnnn0.53
8nnnn0.83
9nnnn−0.72
a Predicted by DataWarrior algorithm; M: mutagenic; T: tumorigenic; I: irritant; R: reproductive effect. Levels—n: none; h: high.
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

Maldonado, J.; Oliva, A.; Guzmán, L.; Molinari, A.; Acevedo, W. Synthesis, Anticancer Activity, and Docking Studies of Novel Hydroquinone-Chalcone-Pyrazoline Hybrid Derivatives. Int. J. Mol. Sci. 2024, 25, 7281. https://doi.org/10.3390/ijms25137281

AMA Style

Maldonado J, Oliva A, Guzmán L, Molinari A, Acevedo W. Synthesis, Anticancer Activity, and Docking Studies of Novel Hydroquinone-Chalcone-Pyrazoline Hybrid Derivatives. International Journal of Molecular Sciences. 2024; 25(13):7281. https://doi.org/10.3390/ijms25137281

Chicago/Turabian Style

Maldonado, Javier, Alfonso Oliva, Leda Guzmán, Aurora Molinari, and Waldo Acevedo. 2024. "Synthesis, Anticancer Activity, and Docking Studies of Novel Hydroquinone-Chalcone-Pyrazoline Hybrid Derivatives" International Journal of Molecular Sciences 25, no. 13: 7281. https://doi.org/10.3390/ijms25137281

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