Next Article in Journal
Size-Dependent Antibacterial, Antidiabetic, and Toxicity of Silver Nanoparticles Synthesized Using Solvent Extraction of Rosa indica L. Petals
Next Article in Special Issue
Achillea fragrantissima (Forssk.) Sch.Bip Flower Dichloromethane Extract Exerts Anti-Proliferative and Pro-Apoptotic Properties in Human Triple-Negative Breast Cancer (MDA-MB-231) Cells: In Vitro and In Silico Studies
Previous Article in Journal
In Silico Antiprotozoal Evaluation of 1,4-Naphthoquinone Derivatives against Chagas and Leishmaniasis Diseases Using QSAR, Molecular Docking, and ADME Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dihydropyrazole-Carbohydrazide Derivatives with Dual Activity as Antioxidant and Anti-Proliferative Drugs on Breast Cancer Targeting the HDAC6

by
Irving Balbuena-Rebolledo
1,2,3,
Astrid M. Rivera-Antonio
1,2,
Yudibeth Sixto-López
3,4,
José Correa-Basurto
3,*,
Martha C. Rosales-Hernández
2,
Jessica Elena Mendieta-Wejebe
2,
Francisco J. Martínez-Martínez
5,
Ivonne María Olivares-Corichi
6,
José Rubén García-Sánchez
6,
Juan Alberto Guevara-Salazar
7,
Martiniano Bello
3 and
Itzia I. Padilla-Martínez
1,*
1
Laboratorio de Química Supramolecular y Nanociencias, Unidad Profesional Interdisciplinaria de Biotecnología, Instituto Politécnico Nacional, Avenida Acueducto s/n, Barrio la Laguna Ticomán, Ciudad de México 07340, Mexico
2
Laboratorio de Biofísica y Biocatálisis, Sección de Estudios de Posgrado e Investigación, Escuela Superior de Medicina, Instituto Politécnico Nacional, Plan de San Luis y Salvador Díaz Mirón s/n, Casco de Santo Tomas, Ciudad de México 11340, Mexico
3
Laboratorio de Diseño y Desarrollo de Nuevos Fármacos e Innovación Biotecnológica, Escuela Superior de Medicina, Instituto Politécnico Nacional, Plan de San Luis y Díaz Mirón, s/n, Col. Casco de Santo Tomas, Ciudad de México 11340, Mexico
4
Departamento de Química Farmacéutica y Orgánica, Facultad de Farmacia, Universidad de Granada, Campus de Cartuja, 18071 Granada, Spain
5
Facultad de Ciencias Químicas, Universidad de Colima, Km. 9 Carretera Colima-Coquimatlán, C.P. Coquimatlán, Colima 28400, Mexico
6
Laboratorio de Oncología Molecular y Estrés Oxidativo de la Escuela Superior de Medicina, Instituto Politécnico Nacional, Plan de San Luis y Díaz Mirón, s/n, Col. Casco de Santo Tomas, Ciudad de México 11340, Mexico
7
Departamento de Farmacología, Escuela Superior de Medicina, Instituto Politécnico Nacional, Plan de San Luis y Díaz Mirón, S/N, Ciudad de México 11340, Mexico
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2022, 15(6), 690; https://doi.org/10.3390/ph15060690
Submission received: 2 April 2022 / Revised: 20 May 2022 / Accepted: 26 May 2022 / Published: 31 May 2022
(This article belongs to the Special Issue Drug Candidates for the Treatment of Breast Cancer)

Abstract

:
Breast cancer (BC) is the most frequently diagnosed cancer and is the second-most common cause of death in women worldwide. Because of this, the search for new drugs and targeted therapy to treat BC is an urgent and global need. Histone deacetylase 6 (HDAC6) is a promising anti-BC drug target associated with its development and progression. In the present work, the design and synthesis of a new family of dihydropyrazole-carbohydrazide derivatives (DPCH) derivatives focused on HDAC6 inhibitory activity is presented. Computational chemistry approaches were employed to rationalize the design and evaluate their physicochemical and toxic-biological properties. The new family of nine DPCH was synthesized and characterized. Compounds exhibited optimal physicochemical and toxicobiological properties for potential application as drugs to be used in humans. The in silico studies showed that compounds with –Br, –Cl, and –OH substituents had good affinity with the catalytic domain 2 of HDAC6 like the reference compounds. Nine DPCH derivatives were assayed on MCF-7 and MDA-MB-231 BC cell lines, showing antiproliferative activity with IC50 at μM range. Compound 2b showed, in vitro, an IC50 value of 12 ± 3 µM on human HDAC6. The antioxidant activity of DPCH derivatives showed that all the compounds exhibit antioxidant activity similar to that of ascorbic acid. In conclusion, the DPCH derivatives are promising drugs with therapeutic potential for the epigenetic treatment of BC, with low cytotoxicity towards healthy cells and important antioxidant activity.

Graphical Abstract

1. Introduction

According to the World Health Organization, breast cancer (BC) is the leading cause of women’s death, with greater than two million new cases and more than six hundred thousand deaths per year worldwide [1,2]. BC is considered a heterogeneous disease and can be classified as: luminal A, luminal B, or triple-negative breast cancer (TNBC). Luminal A is characterized by the presence of estrogenic receptor (ER+), presence or absence of progesterone receptor (PR±), and absence of epidermal growth factor receptor-2 (HER2-); luminal B is characterized by (ER+), (PR±), and (HER2+); whereas the TNBC lacks expression of the three receptors (ER-), (PR-) and (HER2-) [3,4]. Several BC subtypes are treated with different primary therapeutic protocols, but none of them use epigenetic drugs [5]. Interest in treating BC with epigenetic drugs is increasing, mostly because epigenetic alterations such as DNA methylation and acetylation status of histones have been identified as important factors that contribute to tumorigenesis and BC progression [6,7]. Currently, there are no efficient, targeted treatment options for TNBC. Therefore, the identification of new drug candidates for this treatment is an emergent field of research.
Protein acetylation balance is regulated by histone acetyltransferases (HAT) and deacetylases (HDACs); these enzymes play an essential role in post-translational modifications [8,9]. HDACs are a family of hydrolases that remove acetyl groups of lysine residues from histones [10,11] and regulate the expression of tumor suppressor genes, cell cycle progression, and epigenetic transcription [12,13,14]. Cancer, autoimmune, and psychiatric diseases are some human diseases associated with HDACs malfunctioning [15,16,17]. There are eighteen HDAC isoforms identified in mammals, classified into four classes: class I (HDAC 1, 2, 3, and 8) [18]; class II, which is subdivided into IIa (HDAC 4, 5, 7, and 9) and IIb (HDAC 6 and 10); class III, which are also called sirtuins (HDAC 12-18); and class IV (HDAC11) [19]. Classes I, II, and IV are Zn2+-dependent, while class III are NAD+-dependent [20,21,22]. HDAC6 is primarily expressed in the cytoplasm and encodes a protein of 1215 amino acid residues, the most essential protein of the HDACs family [23]. HDAC6 has a particular structure; it is the only HDAC that contains an internal dimer of two functional catalytic domains, which are named DD1 (G87-G404) and DD2 (G482-G800). They are located at the N-terminal and the central region of the protein-bound by the linker region (D405-T481), where dynein motor binding (DMB) domain is found (V439-V503). HDAC6 maintains the acetylation balance of a wide variety of cytoplasmic proteins [19,24,25,26,27], and its overexpression has been associated with various leading diseases such as cancer [28], neurodegenerative diseases, and pathological autoimmune response. For these reasons, selective HDAC6 inhibitors have been extensively investigated to treat these diseases [29,30,31,32,33,34]. Even when HDAC6 possesses two catalytic domains, its activity can only be attributed to DD2. Indeed, entire HDAC6 inhibitors have been developed to date target this domain, including the selective inhibitor tubacin [31,35,36].
On the other hand, oxidative stress leads to several chronic degenerative diseases and disorders, including cancer [37]. BC etiology is multifactorial; moreover, it has been clearly linked to oxidative stress as an essential risk factor [38]. The oxidative stress induced by reactive oxygen species (ROS) is considered as a dynamic imbalance between endogenous levels of antioxidants and the amount of antioxidants lost to ROS scavenging and protects against their harmful effects [39]. In this context, some studies support that antioxidant supplements may reduce the risk of BC recurrence or BC-related mortality [40,41] more than exerting a protective effect [42]. Within tumor cells, increased ROS levels create an inflammatory environment conducive to tumor progression and dissemination to distant organs [43]. Therefore, attenuation of oxidative stress with an antioxidant should result in reduced size and likelihood of metastasis. In this sense, several efforts have been made to combine anti-inflammatory [44] and antioxidant [45,46] effects in the same molecule or in mixtures [47] for potential anticancer treatments.
Therefore, attention to the development of novel and effective anticancer agents with more selectivity and fewer associated side effects, is required for the disease’s eradication. In this context, 4,5-dihydropyrazole derivatives have attracted attention due to their biological activities, such as anti-inflammatory, antidepressant, and potent antiproliferative activity, specifically against BC [48,49,50]. Particularly relevant is N-(4-hydroxybenzyl)-1,3,4-triphenyl-4,5-dihydro-1H-pyrazole-5-carboxamide (BHX), Figure 1, whose activity against cancer as a Wnt/β-catenin-signaling inhibitor has been demonstrated [51].
In this work, a set of nine dihydropyrazole-carbohydrazide derivatives (DPCH) with potential inhibitory activity on the DD2-HDAC6 domain were modelled in silico. They were synthesized and evaluated in their physicochemical and toxicobiological properties. Moreover, docking simulations on the DD2-HDAC6 domain were performed in order to obtain the non-bonding interactions and the binding free energy values (ΔG°). The results were compared with tubacin, trichostatin A (TSA), and suberoylanilide hydroxamic acid (SAHA) as reference compounds, Figure 1. Antiproliferative assays on BC cell lines MCF-7 and MDA-MB-231 and the nonmalignant cells 3T3/NIH (fibroblast cells) and MCF10A (breast epithelial cells) were performed. MCF-7 and MDA-MB-231 cell lines were considered as models of the most common and the most aggressive subtypes of BC, respectively. Additionally, the overexpression of HDAC6 in both BC cell lines has been demonstrated [28,52]. MCF-7 is the A luminal type, ER and PR (+); MDA-MB-231 is the C-type (claudin-low) and triple-negative—ER, PR and HER2 (−)—also known as triple negative BC (TNBC). Finally, the in vitro inhibition of HDAC6 enzyme as well as its antioxidant properties was demonstrated. The results were supported by quantitative structure–activity relationship analysis (QSAR).

2. Results and Discussion

2.1. Design Features of Compounds 2ai

Several studies have suggested that the catalytic tunnel of DD2-HDAC6 is wider and shallower than other isoforms [53]. Thus, the inclusion of large, bulky aromatic rings in the designed molecules could be useful in fitting into the cap region of the enzyme, more specifically into pockets L1 and L2 [54]. In fact, the aromatic rings are useful for increasing affinity and selectivity by DD2-HDAC6 over the other isoforms, which are mediated by noncovalent interactions [53,55,56]. On the other hand, the hydroxamic acid is the most extended chelating zinc group used in HDACs inhibitors, even though this group is susceptible to be degraded and it is not stable in the organism [57,58]. Additionally, several researchers have focused on replacing the hydroxamic group for other functionalities [54]. In this context, we provide a set of novel compounds based on 4,5-dihydropyrazole heterocycle with a pending 4-carbohydrazide group and bearing several aromatic rings, Figure 2. Despite the bulky effect contribution, they are expected to bind to the cap region of DD2-HDAC6 domain through hydrogen bonding and π-interactions.

2.2. Synthesis of DPCH Derivatives

The synthesis of DPCH derivatives (2ai) is depicted in Scheme 1. This approach includes two reactions. Firstly, a Knoevenagel condensation between the substituted salicylaldehyde and ethyl 3-oxo-3-phenylpropanoate was achieved to yield the corresponding 3-benzoyl-2H-1-benzopyran-2-one 1ai. In the second step, compounds 1ai were treated with phenylhydrazine and glacial acetic acid as a catalyst under reflux of EtOH to obtain the corresponding DPCH derivatives 2ai with poor-to-good yields (20–60%) as a racemic mixture. The compounds precipitated, leaving in solution the more soluble 1,3-diphenylchromene[4,3-c]pyrazole-4(1H)-ones and their corresponding phenylhydrazones [59]. It is worth highlighting that the final products required no further purification (purity > 98%). At this point, a brief comparison with BHX synthesis, a closely related compound to 2ai, seems appropriate, Figure 1. The synthesis of BHX is attained after four steps, with yields after chromatography of 98.9, 10.4, 72.0, and 56.5%, respectively, to give a final 4.2% overall yield [51]. This yield highlights the benefits of the method herein reported.
One of the striking structural features of compounds 2ai is the cis disposition between H4 and H5. This stereochemistry was suggested by the coupling constant value between these hydrogen atoms (3J), which is around 12 Hz. A nuclear Overhauser effect (nOe) experiment was performed for the assignment of the 1H signals belonging to compound 2a, Figure S1. The selected signals were H4, H5, and H21, which are at lower frequencies in the spectrum and separated from each other. In a nOe spectra, all signals are vanished except those corresponding to the hydrogen atoms that are coupled or close in space to the irradiated signal. Then, nOe on H4 (d, δ 4.95) allowed assignment of the amide proton (δ 9.79) and H11 (δ 7.81) and irradiation on H5 allowed assignment of H7 (δ 6.92), whereas the absence of nOe on H19 indicates that the phenolic ring is out of the plane of the pyrazole ring and opposite to H5. Finally, the nOe on H21 (δ 6.03) allowed the NHPh to be assigned at δ 7.56. Once the signals for the ortho protons of the three monosubstituted rings had been identified, the other signals were assigned with homo- and heteronuclear two-dimensional spectra.
The reaction proceeded through the intramolecular 1,4-addition of N1H to the α,β-unsaturated lactone carbonyl of the 3-benzoyl-coumarin-phenyl hydrazone A to form the pyrazole ring. The pyranol ring in B adopts a boatlike conformation with cyclic oxygen and C5 atoms positioned on the vertexes out of the plane of the boat conformer. The enol form B is then tautomerized to the keto form C to give the cis isomer. The stereoselectivity of this reaction is explained because of the steric effect exerted by both the coumarin benzofused ring and the C3-Ph ring that limit the approaching of H+ to the opposite face occupied by H5, leading to the formation of the cis isomer as a single diastereoisomer. The final product 2 is achieved as a racemic mixture after the amidation and ring opening of the pyrone ring of the intermediate C by a second phenylhydrazine molecule, Scheme 2.

2.3. Molecular Structure of Compound 2a

The structure of compound 2a was confirmed by single-crystal X-ray diffraction; it crystallized in the monoclinic crystal system and P21/c space group as the DMSO solvate. The molecular structure is shown in Figure 3; the bond lengths, bond angles, and torsion angles are listed in Table S1. The DPCH ring adopts an envelope conformation with C5 positioned in the vertex, as is revealed by the torsion angles’ values: C4—C3—N2—N1, -0.7(3)°; C3—C4—C5—N1, 14.20(19)°; and N2—N1—C5—C4, -12.7(2)°. This conformation is the most frequently observed in the 22 hits of dihydro-pyrazole scaffold retrieved from the CCDC [60]. The torsion angle of C24—N3—N4—C20 of 113.6(3)°, in the phenyl-hydrazone fragment, is closer than that observed in a similar compound (146.3(3)°) (CCDC-188945) [61]. Additionally, the cis disposition between H4 and H5 is confirmed by C24—C4—C5—C14’s torsion angle value of −21.0(3)°. The supramolecular architecture is given by the following hydrogen bonding interactions (D—H∙∙∙A): N3—H3∙∙∙O24, N4—H4···O25 (DMSO), O15—H15∙∙∙O25 (DMSO), and C4—H4A∙∙∙O24. The geometric parameters of these interactions are listed in Table S2.

2.4. In Silico Studies

2.4.1. Docking Simulation

The affinity of DD2-HDAC6 enzyme towards 2ai was theoretically studied using molecular docking simulations. Docking studies allowed us to obtain the free energy of ligand–receptor binding (ΔGb°) as well as the dissociative equilibrium constant Kd of the nine 2a–i-DD2-HDAC6 complexes and reference compounds (tubacin, TSA, and SAHA). Analysis of the two enantiomers (4S, 5S) and (4R, 5R) of 2ai present in the racemic mixture of reaction was performed to elucidate the effect of the configuration on ΔGb°. Molecular docking results are listed in Table S3. According to the docking calculations, all compounds are active towards DD2-HDAC6. ΔGb° values are in the −7.78 to −6.85 kcal/mol range, close to the values obtained for the reference compounds—TSA had a value of −8.59 kcal/mol and SAHA had a value of −7.02 kcal/mol—but 3.1–2.2 kcal/mol smaller than the ΔGb° value of tubacin of −9.97 kcal/mol. Moreover, the ΔGb° difference between the (4S, 5S) and the (4R, 5R) enantiomers is small—from 0.61 to −0.01 kcal/mol—and in most cases is in favor of the first enantiomer. Therefore, further in silico calculations were performed only on the (4S, 5S) enantiomer.
The calculated binding modes of 2ai-DD2-HDAC6 and tubacin-DD2-HDAC6 complexes show that all ligands reached the catalytic binding site of DD2-HDAC6. Compounds 2ai are anchored on the surface binding domain, and one of the four aromatic rings is slipped into the hydrophobic channel. Those compounds bearing a substituent in the para position relative to the phenolic group (2b, 2c, 2g, and 2h; not 2d) seem to favor the positioning of the C5-PhOH ring into the hydrophobic channel over those ortho-substituted (2e and 2i), meta-substituted (2f), or unsubstituted (2a), which favor C3-Ph or N-Ph insertion into the hydrophobic channel, respectively, Figure 4. CONHNHPh residue contributes through NH∙∙∙N, NH∙∙∙O, and OH∙∙∙O hydrogen bonding to the ligand anchorage into the rim of the DD2 domain. All ligands interact with S568, F620, F680, H651, F679, and L749 amino acid residues (AAR) of DD2-HDAC6—the same as tubacin, TSA, and SAHA, the compounds used as reference—as well as with H611 (67% of incidence), G619 (72%), D567, T678, and Y782 (33% each), Table S4, through hydrogen bonding, electrostatic, π−π type, and mostly hydrophobic interactions, Figure 4. It is worth mentioning that these interactions are common with other recently reported HDAC6 inhibitors [62,63,64]. As can be seen, compounds 2ai are locking the entrance to the catalytic tunnel by the 4,5-dihydro-pyrazole moiety, effectively guarding the active site of HDAC6, Figure 4. This binding mode is similar to that shown by tubacin, whose HDAC6 inhibitory activity has been attributed to its bulky and relatively complex capping group [65]. The complete set of binding conformation of complexes between compounds 2ai and monomeric DD2-HDAC6 is displayed in Figure S2.

2.4.2. Evaluation of Physicochemical and Toxicobiological Properties

The theoretical physicochemical and toxicobiological properties of compounds 2ai, tubacin, TSA, and SAHA (the triad of compounds used as reference) were analyzed trough Osiris DataWarrior and Osiris Property Explorer software, respectively. The results are listed in Table 1, where it can be seen that most of the tested compounds satisfy Lipinski’s five rules [66]. In general, narrow intervals were observed for MW (448.52–557.44 g mol−1), logP (2.2–3.4), number of hydrogen acceptors (HA = 6–7), and hydrogen donors (HD = 3–4), as well as for the number of rotatable bonds (RB = 6–7). Particularly, compounds 2d and 2h, substituted with bromine, are out of the range for optimal MW. These theoretical predictions are of high importance for the analysis of hundreds of drugs. Many of them are approved by the U.S. Food and Drug Administration (FDA), even when they violated more than two of Lipinski’s rules [67]. Additionally, the LogS, topological polar surface area (TPSA = 77.0–97.2 Å2), absorption percentage by passive diffusion (%ABS = 109 ± 0.345–TPSA), and molar refractivity (MR = 132.1–147.0 cm3 mol−1) [68] values of each compound were obtained [69]. The whole set of DPCH derivatives is predicted to have good lipid membrane absorption, with %ABS values in the 76–82% range. These values are similar to TSA (85%) and SAHA (82%) and are much better than those for tubacin (51%). Finally, good toxicobiological properties were predicted for 2ai, Table S5, except for the high risk of being as tumorigenic as TSA. However, it is very common that treatments used for cancer have a critical toxicity profile and cause a number of side effects [70]. Even though substantial progress has been made in antitumor drugs, drug resistance and high toxicity still limit their clinical application [71,72,73].

2.5. In Vitro Pharmacological Evaluation

2.5.1. Cell Viability Assays

The antiproliferative activities of 2ai were evaluated on two types of BC cancer cell lines (MCF-7 and MDA-MB-231). The cytotoxic evaluation of 2ai was conducted by MTT assay in MCF-7 cells (Figure S3), MDA-MB-231 (Figure S4), the nontumorigenic 3T3/NIH cells lines (Figures S5 and S6), and the nonmalignant breast epithelial cells MCF10A (Figures S7 and S8) and compared with SAHA and TSA as reference drugs. The IC50 for SAHA was similar to that reported elsewhere [74]. Results showed that the cytotoxicities of 2ai are dose-dependent toward both BC cell lines, with IC50 values in the µM range, Table 2. The compounds with the best antiproliferative activity in MCF-7 cells were 2c–g (IC50 = 23–28 µM), whereas those with the best activity in MDA-MB-231 cells were 2b and 2d (IC50 = 24–26 µM), followed by 2c and 2e–g (IC50 = 32–33 µM). Moreover, the unsubstituted compound 2a and compounds 2c and 2i, substituted with –OR (R = Me, Et) group, were less cytotoxic to normal 3T3/NIH cells (IC50 > 100 µM) than the rest of the compounds, particularly those substituted with an –OH group (2e–g). Although the tested compounds were not better than TSA and SAHA, they exhibited cytotoxic activities similar to pyrimethamine-hydroxamic acid derivatives towards MCF-7 and MDA-MB-231 cell lines [74]. However, they were better than pyrrolo[2,3-d]pyrimidine-based HDAC inhibitors in MDA-MB-231 cells [75]. In the case of compound 2c, the IC50 value in MDA-MB-231 cells was slightly larger (33 ± 1 μM) than the value of BHX (19.3 μM) but less cytotoxic to nonmalignant MCF-10A (>100μM) than BHX (31.06 μM) [50]. The last comparison allows us to conclude that the cis disposition between the C5-Ph and 4-CONHNHPh groups, in contrast to the trans disposition between C5-Ph and C4-Ph in BHX, could be related with of the lower cytotoxicity of compounds 2c, 2h, and 2i against nontumorigenic cells compared to BHX.

2.5.2. Wound Closure Assays in the MDA-MB-231 BC Line

MDA-MB-231 cells are a very aggressive and metastatic cancer line that tends to migrate to other organs. It is known that overexpression of HDAC isoforms (1, 4, 6, and 8) in both MDA-MB-231 and MCF-7 cells is responsible of the invasiveness and migration capabilities of human breast cancer cells [52]. Compounds 2b and 2c at ½-IC50 concentration (15 µM) were assayed to establish their capability to prevent cell migration. The percent of wound closure was measured after 16, 24, and 48 h of being inflicted. Figure 5a–c shows that the wound closure begins at 24 h, reaching more than 80% after 48 h in the control, whereas treated cells were less than 20% after 48 h of incubation. This result confirms the capability of compounds 2b and 2c to prevent the migration of MDA-MB-231 cells.

2.5.3. In Vitro HDAC6 Human Recombinant Inhibition

Compound 2b was selected to perform the HDAC6 inhibition assay because it showed the best ΔGb of the 2b-DD2-HDAC6 complex of -7.86 kcal mol−1. The assay was validated using TSA as positive control, finding a Ki similar to the reported value [76]. Compound 2b inhibited HDAC6 in a dose-dependent manner at IC50 = 12 ± 3 µM, Figure 5d. Although the IC50 value is higher than the reported for other HDAC6 inhibitors (nM) [64,74,77], this result could be explained due to the absence of a N-OH group, which is capable of chelating the Zn2+ present in the HDAC6 catalytic site. The capability of reaching the cytoplasm, where the HDAC6 is overexpressed [78], was investigated. The intracellular location of compound 2b was confirmed with confocal laser microscopy taking advantage of the fluorescence exhibited by this compound. The MCF-7 cells were exposed to 2b at 10 µM for 30 min. Figure 5e is a live cell imaging that shows the entrance of the compound (blue) into the cell. This result agrees with those obtained from the in silico study since, according to the physicochemical properties determined by the Lipinski’s rules, the compounds presented optimal properties for crossing the cell membrane.

2.6. QSAR Analysis

The correlations found through a QSAR analysis of DPCH derivatives on the most common (MCF-7) and most aggressive (MDA-MB-231) BC cell lines are described below. The pIC50 values of the DPCH derivatives show a parabolic correlation with the Es descriptor proposed by Taft [79], in the MCF-7 cell line. Therefore, the inhibitory activity on proliferation is dependent on the size of the molecules, the relationship shows that derivatives with MR values between 134 and 139 cm3 mol−1 have the best activities (2cg), while smaller or larger derivatives are significantly less active, Figure 6. Likewise, a similar correlation was found for DPCH derivatives on the MDA-MB-231 cell line with the steric descriptor (Es). However, the relationship between the molecular size and antiproliferative activity is more evident; that is, derivatives with medium sizes (2b, 2c, 2d) showed the best activity in relation to those compounds of smaller (2ag) or larger size (2h, 2i), Figure 6, see Figure S9 for 3T3/NIH.
On the other hand, the pIC50 values of the DPCH derivatives showed parabolic correlations with the liposolubility descriptor proposed by Hansch (π) [80] and the electronic descriptor proposed by Hammett (σH) [81] on the 3T3/NIH and MCF-10A cell lines. The less toxic derivatives with both cell lines lie at the minimum of the curves. These have π values between −0.1 and 0.3, corresponding to logP values between 2.6 and 3.0 (2a, 2c, 2i and 2h), and small, negative values for σ (2a and 2c). It is worth mentioning that in the case of the σ descriptor, only those derivatives with substitutions in the meta and para positions were considered since σ cannot be appropriately estimated in the ortho positions due to the overlap with steric effects.

2.7. Antioxidant Activity

The radical scavenging activity (RSA) of 2ai was assessed by the DPPH test, using ascorbic acid (AA) as control. This assay is widely used to evaluate the antioxidant capabilities of natural and synthetic compounds, where DPPH is the free radical which can accept an electron or hydrogen atom and become reduced [82]. The results are shown in Figure 7a; all compounds exhibited antioxidant activities above 75% at 100 µM, with 2c, 2e, 2h, and 2i being the most active (up to 90%) and similar to AA (93%). The DPPH RSA IC50 values of all compounds are in the 16–38 μM range; the best performances are shown by compounds 2c (16 ± 4 μM) and 2h (17 ± 3 μM) with values comparable to AA (13 ± 2 μM, 13.9 μM) [83]. The complete dose–response curves are displayed in Figure S10. It is worth noting that small IC50 values are desired for antioxidant and antiproliferative activities against malignant cells (MCF-7 and MDA-MB231) and large IC50 values for antiproliferative activity against nonmalignant cells (MCF-10A). This relationship is clearly appreciated for compounds 2c, 2a, and 2i in the spiderweb chart shown in Figure 7b. Therefore, these compounds can be considered as effective dual anticancer–antioxidant agents with reduced cytotoxicity in normal breast cells by decreasing ROS production—characteristics desired for diminishing some undesirable side effects of chemotherapy [47]—and also for BC treatment [84].

3. Materials and Methods

3.1. Instrumental and Chemicals

All reagents and solvents were purchased from commercial suppliers and used without further purification. Thin-layer chromatography (TLC) developments were performed on silica gel coated (Merck 60 F254) aluminum foils. Yields are reported after final isolated products with 98–100% purity (HPLC-Agilent Technologies 1260 Infinity Series system). Melting points were measured in an Electrothermal IA 91000 devise and are uncorrected. Proton and carbon-13 nuclear magnetic resonance (1H and 13C NMR) spectra were recorded on a Varian Mercury NMR spectrometer operating at 300 MHz (1H, 300.08; 13C, 75.46 MHz), using deuterated dimethylsulfoxide (DMSO-d6) as a solvent; chemical shift values (δ) are reported in parts per million (ppm), using as reference the residual solvent peaks (1H, δ 2.50; 13C, δ 39.52) and coupling constants nJ(H–H) are in Hz. Multiplicity of the signals are expressed as: s (singlet), d (doublet), t (triplet), q (quartet), or m (multiplet), Figures S11–S28. The complete assignment of 1H and 13C-NMR were performed with COSY and HETCOR 2D experiments, Figures S29 and S30, and nOe 1D spectra, Figure S1. The numbering scheme for NMR assignments is shown in Figure 3a. Infrared (IR) spectra were recorded neat in a Perkin-Elmer Spectrum GX series with an FT-IR System Spectrophotometer using the ATR devise, the intensity of the signals was indicated as: weak (w), medium (m), strong (s), or very strong (vs), Figures S31–S39. Mass spectrometry was performed on an Agilent UHPLC-Mass Spectrometer 6545 Q-TOF LC/MS, using acetonitrile as solvent, Figures S40–S53, for purity.

3.2. X-ray Structure Determination

General crystallographic data for 2a has been deposited in the Cambridge Crystallographic Data Centre as supplementary publication number CCDC 2108339. A summary of the collection and refinement of the X-ray data is listed in Table S6. Single crystal X-ray diffraction data were collected on an Oxford Xcalibur Ruby Gemini area detector diffractometer at 293(2) K with Mo Kα radiation (λ = 0.71073 Å). Cell refinement and data reduction were carried out with the CrysAlis RED software [85]. The structures were solved by direct methods using the SHELXS2014 program [86] of the WINGX package [87]. The final refinement was performed by full-matrix least-squares methods using the SHELX2014 program [86]. H atoms on C were positioned geometrically and treated as riding atoms with C−H 0.93−0.98 Å, Uiso(H) = 1.2 eq(C), and H atoms on O or N were found by Fourier difference and freely refined. Platon [88] and Mercury [89] were used to prepare the material for publication.

3.3. Chemical Synthesis of Substituted DPCH Derivatives 2ai

Compounds 1ai are known, but they are not commercially available. Therefore, they were synthesized as follows: in a 250 mL ball flask, the corresponding amount of salicylaldehyde was placed together with ethyl benzoyl acetate in 1:1.1 molar ratio in 30 mL of ethyl alcohol as solvent and 3 drops of piperidine as catalyst. It was allowed to stir at reflux for 12 h. The product was filtered under vacuum and washed with ethanol. Their spectroscopic characterization agrees with the literature [90,91,92] (see ESI).
(±)-5-(2′-Hydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2a). To a solution of 0.300 g (1.20 mmol) of 1a dissolved in 25 mL of ethanol, 0.35 mL of phenylhydrazine (3.6 mmol), 4.0 mL of distilled water and 4 drops of glacial acetic acid were added. The reaction was allowed to reflux with magnetic stirring for 24 h. After completion, the reaction mixture was allowed to cool at room temperature or until the formation of a white solid was observed. The solid was filtered under vacuum, washed with ethanol (2 × 3 mL) and allowed to dry at room temperature to obtain 0.158 g (0.35 mmol, 30% yield, 98.92% purity) of a white fluorescent solid, mp = 207–210 °C. IR (cm−1): 3345 (w), 3253 (w) (N-H, O-H), 1645 (m, C=O), 1594 (m), 1494 (s) 1455 (m) (C=C, Ph), 1365 (m), 1223 (m), 772, 750, 688 (vs, C-H Aromatic out of plane). RMN 1H δ: 10.03 (s, 1H, OH), 9.79 (s, 1H, CONH), 7.81 (dd, 2H, 3J = 8.2, 4J = 1.8, H11), 7.56 (s, 1H, PhNH), 7.43 (dd, 2H, 3J = 7.6, 4J = 8.2, H12), 7.37 (dd, 1H, 3J = 7.6, 4J = 1.8, H13), 7.14 (dd, 3H, 3J = 7.3, 3J = 8.2, H8), 7.14 (t, 3J = 7.6, H17), 7.00 (d, 2H, 3J = 7.6, H16, H19), 6.92 (d, 2H, 3J = 8.2, H7), 6.86 (dd, 2H, 3J = 8.2, 3J = 7.6, H22), 6.77 (t, 1H, 3J = 7.3, H9), 6.57 (t, 1H, 3J = 7.6, H18), 6.53 (t, 1H, 3J = 7.6, H23), 6.03 (d, 2H, 3J = 8.2, H21), 5.62 (d, 1H, 3J = 12, H5), 4.95 (d, 1H, 3J = 12, H4). RMN 13C δ: 167.6 (CO), 154.9 (C15), 148.9 (C3), 147.7 (C20), 145.8 (C6), 132.4 (C10), 129.8 (C19), 129.2 (C13), 129.1 (C12), 129.0 (C8), 128.9 (C17), 128.8 (C22), 126.2 (C11), 122.2 (C14), 120.2 (C9), 119.4 (C18), 118.4 (C23), 115.3 (C16), 115.0 (C7), 112.2 (C21), 63.1 (C5), 54.7 (C4). Mass analysis [M-H]+ (m/z): 449.1978 found, 449.1978 calculated.
(±)-5-(5′-Chloro-2′-hydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2b). Synthesized as described for 2a starting from 0.300 g (1.05 mmol) of 1b, 0.35 mL of phenylhydrazine (3.6 mmol) to obtain 0.170 g (0.35 mmol, 33% yield, 97.88% purity) of a white fluorescent solid, mp = 205–207 ºC. IR (cm−1): 3330 (w), 3257 (br) (N-H, O-H), 1645 (m, C=O), 1595 (m), 1494 (s) 1419 (m) (C=C, Ph), 1362 (m), 1275 (m), 809 (m, C-Cl), 753, 692 (vs, C-H Aromatic out of plane). 1H NMR δ: 10.43 (s, 1H, OH), 9.88 (s, 1H, CONH), 7.79 (d, 2H, 3J = 8.2, H11), 7.57 (s, 1H, PhNH), 7.40 (m, 3H, H12, H13), 7.19 (d, 1H, 3J = 8.8, H17), 7.18 (dd, 2H, 3J = 7.6, 3J = 8.2, H8), 7.01 (d, 1H, 3J = 8.8, H16), 6.93 (s, 1H, H19), 6.90 (d, 2H, 3J = 8.2, H7), 6.89 (t, 2H, 3J = 7.6, H22), 6.81 (t, 1H, 3J = 7.6, H9), 6.56 (t, 1H, 3J = 7.6, H23), 6.10 (d, 2H, 3J = 7.6, H21), 5.77 (d, 1H, 3J = 12.3, H5), 4.97 (d, 1H, 3J = 12.3, H4). RMN 13C δ: 167.3 (CO), 154.0 (C15), 148.8 (C3), 148.0 (C20), 145.6 (C6), 132.2 (C10), 129.4 (C13), 129.2 (C8), 129.1 (C12), 128.9 (C17), 128.8 (C22), 126.3 (C11), 124.5 (C18), 122.9 (C14), 120.5 (C9), 118.6 (C23), 117.1 (C16), 114.9 (C7, C19), 112.2 (C21), 62.8 (C5), 54.7 (C4). Mass analysis [M-H]+ (m/z): 483.1587 found, 483.1588 calculated.
(±)-5-(2′-Hydroxy-5′-methoxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2c). Synthesized as described for 2a starting from 0.300 g (1.07 mmol) of 1c, 0.35 mL of phenylhydrazine (3.6 mmol) to obtain 0.32 g (0.66 mmol, 63% yield, 98.86% purity) of a white fluorescent solid, mp = 205–206 °C. IR (cm−1): 3452 (w), 3306 (w), 3225 (w) (N-H, O-H), 1654 (m, C=O), 1595 (m), 1494 (s), 1446 (m) (C=C, Ph), 1427 (m, CH3), 1366 (m), 1220 (m), 773 (vs), 750 (s), 689 (vs) (C-H Aromatic out of plane). 1H NMR δ: 9.80 (d, 1H, 3J = 1.8, CONH), 9.64 (s, 1H, OH), 7.82 (d, 2H, 3J = 8.2, H11), 7.80 (s, 1H, PhNH), 7.45 (dd, 2H, 3J = 7.6, 3J = 8.2, H12), 7.40 (t, 1H, 3J = 7.6, H13), 7.18 (dd, 2H, 3J = 7.6, 3J = 8.2, H8), 6.90 (d, 2H, 3J = 8.2, H7), 6.89 (t, 3H, 3J = 7.6, H22), 6.82 (d, 1H, 3J = 7.6, H9), 6.81 (d, 1H, 3J = 7.6, H16), 6.78 (dd, 1H, 3J = 7.6, 4J = 3.0, H17), 6.60 (d, 1H, 4J = 3.0, H19), 6.57 (t, 1H, 3J = 7.6, H23), 6.08 (d, 1H, 3J = 7.6, H21), 5.60 (d, 1H, 3J = 11.7, H5), 4.95 (d, 1H, 3J = 11.7, H4), 3.41 (s, 3H, CH3). 13C NMR δ: 167.3 (CO), 151.9 (C15), 148.6 (C3), 148.5 (C20), 147.5 (C6), 145.5 (C18), 132.0 (C10), 129.0 (C13), 128.82 (C12), 128.77 (C8), 128.5 (C22), 125.9 (C11), 122.9 (C14), 120.0 (C16), 118.1 (C23), 115.6 (C19), 115.4 (C9), 114.7 (C7), 113.4 (C17), 111.9 (C21), 62.9 (C5), 55.1 (OMe) 54.4 (C4). Mass analysis [M-H]+ (m/z): 479.2087 found, 479.2083 calculated.
(±)-5-(5′-Bromo-2′-hydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2d). Synthesized as described for 2a starting from 0.300 g (0.911 mmol) of 1d, 0.35 mL of phenylhydrazine (3.6 mmol) to obtain 0.16 g (0.30 mmol, 33% yield, 98.54% purity) of a white fluorescent solid, mp = 201–203 °C. IR (cm−1): 3338 (w), 3252 (br) (N-H, O-H), 1641 (m, C=O), 1596 (m), 1493 (s) 1415 (m) (C=C, Ph), 1362 (m), 1274 (m), 767 (s), 753, 692 (vs, C-H Aromatic out of plane). RMN 1H δ: 10.50 (s, 1H, CONH), 9.93 (s, 1H, OH), 7.86 (d, 2H, 3J = 7.0, H11), 7.62 (s, 1H, PhNH), 7.48 (t, 2H, 3J = 7.5, H12), 7.46 (t, 1H, 3J = 7.8, H13), 7.36 (dd, 1H, 3J = 8.8, 4J = 2.3, H17), 7.24 (t, 2H, 3J = 7.6, H8), 7.12 (d, 1H, 4J = 2.3, H19), 7.03 (d, 1H, 3J = 8.8, H16), 6.97 (d, 2H, 3J = 7.6, H7), 6.96 (t, 2H, 3J = 7.6, H22), 6.87 (t, 1H, 3J = 7.6, H9), 6.62 (t, 1H, 3J = 7.6, H23), 6.17 (d, 2H, 3J = 7.6, H21), 5.63 (d, 1H, 3J = 11.7, H5), 5.04 (d, 1H, 3J = 11.7, H4). 13C NMR δ: 167.1 (CO), 154.3 (C15), 148.6 (C3), 147.8 (C20), 145.4 (C6), 132.0 (C10), 131.8 (C19), 131.6 (C17), 129.3 (C13), 129.1 (C8), 129.0 (C12), 128.7 (C22), 126.2 (C11), 124.9 (C14), 120.3 (C9), 118.4 (C23), 117.5 (C16), 114.8 (C7), 112.0 (C21), 110.5 (C18), 62.6 (C5), 54.6 (C4). Mass analysis [M-H]+ (m/z): 527.1088 found, 527.1083 calculated.
(±)-5-(2′,3′-Dihydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2e). Synthesized as described for 2a starting from 0.300 g (1.12 mmol) of 1e, 0.5 mL of phenylhydrazine (5.08 mmol), and 5 drops of glacial acetic acid; after 48 h of reaction, 0.075 g (0.16 mmol, 15% yield, 100% purity) of a white fluorescent solid was obtained, mp. = 206–208 °C. IR (cm−1): 3531, 333, 3242 (br) (N-H, O-H), 1649 (m, C=O), 1596 (m), 1494 (s), 1477 (sh), 1366 (m), 1285 (s), 753. 691 (vs, C-H Aromatic out of plane). RMN 1H δ: 9.71 (s, 1H, CONH), 9.54 (s, 1H, OH), 8.89 (s, 1H, OH), 7.79 (d, 2H, 3J = 7.2, H11), 7.49 (s, 1H, PhNH), 7.42 (t, 2H, 3J = 7.2, H12), 7.39 (t, 1H, 3J = 7.1, H13), 7.14 (dd, 2H, 3J = 8.4, 3J = 7.5, H8), 6.93 (d, 2H, 3J = 7.9, H7), 6.90 (t, 2H, 3J = 7.9, H22), 6.77 (t, 1H, 3J = 7.5, H9), 6.74 (d, 1H, 3J = 8.0, H19), 6.53 (t, 1H, 3J = 7.4, H23), 6.50 (d, 1H, 3J = 8.5, H17), 6.38 (d, 1H, 3J = 7.9, H18), 6.08 (d, 2H, 3J = 7.9, H21), 5.63 (d, 1H, 3J = 11.8, H5), 4.93 (d, 1H, 3J = 11.8, H4). RMN 13C δ: 167.5 (CO), 149.0 (C15), 147.6 (C16), 145.9 (C3), 145.3 (C20), 145.2 (C6), 132.5 (C10), 129.2 (C13), 129.1 (C12), 129.0 (C8), 128.9 (C22), 126.2 (C11), 123.1 (C14), 120.1 (C9, C17), 119.2 (C18), 118.4 (C23), 115.0 (C7), 114.8 (C19), 112.3 (C21), 63.3 (C5), 54.7 (C4). Mass analysis [MH+] (m/z): 465.1903 found, 465.1927 calculated.
(±)-5-(2′,4′-Dihydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2f). Synthesized as described for 2a starting from 0.300 g (1.12 mmol) of 1f, 0.35 mL of phenylhydrazine (3.6 mmol), to obtain 0.16 g (0.34 mmol, 30% yield, 98.36% purity) of a white fluorescent solid, mp = 206–208 °C. IR (cm−1): 3241 (br) (N-H, O-H), 1649 (m, C=O), 1606 (sh), 1596 (m) 1494 (s) (C=C, Ph), 1463 (m), 1364 (m), 1217 (m), 772, 752, 692 (vs, C-H Aromatic out of plane). 1H NMR δ: 9.79 (s, 1H, OH), 9.71 (d, 1H, 3J = 1.8, CONH), 9.25 (s, 1H, OH), 7.76 (d, 2H, 3J = 8.2, H11), 7.52 (d, 1H, 3J = 1.8, PhNH), 7.41 (dd, 2H, 3J = 7.1, 3J = 8.2, H12), 7.37 (t, 1H, 3J = 7.1, H13), 7.13 (t, 2H, 3J = 7.6, H8), 6.94 (d, 2H, 3J = 7.6, H7), 6.90 (t, 2H, 3J = 7.6, H22), 6.76 (t, 1H, 3J = 7.6, H9), 6.76 (d, 1H, 3J = 8.2, H19), 6.53 (t, 1H, 3J = 7.6, H23), 6.47 (d, 1H, 3J = 2.3, H16), 6.07 (d, 2H, 3J = 7.6, H21), 6.00 (dd, 1H, 3J = 8.2, 4J = 2.3, H18), 5.52 (d, 1H, 3J = 12.0, H5), 4.83 (d, 1H, 3J = 12.0, H4). 13C NMR δ: 167.8 (CO), 158.3 (C17), 155.8 (C15), 148.9 (C3), 147.5 (C20), 145.9 (C6), 132.6 (C10), 130.4 (C19), 129.11 (C13), 129.01 (C12), 129.0 (C8), 128.7 (C22), 126.1 (C11), 120.0 (C9), 118.4 (C23), 115.0 (C7), 112.6 (C14), 112.3 (C21), 107.2 (C18), 102.5 (C16), 62.9 (C5), 54.7 (C4). Mass analysis [M-H]+ (m/z): 465.1925 found, 465.1927 calculated.
(±)-5-(2′,5′-Dihydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2g). Synthesized as described for 2a starting from 0.300 g (1.12 mmol) of 1g and 0.5 mL of phenylhydrazine (5.08 mmol) to obtain after 48 h of reaction 0.093 g (0.20 mmol, 18% yield, 99.43% purity) of a white fluorescent solid, mp = 206–208 °C. IR (cm−1): 3493 (w), 3302 (br), 3246 (br) (N-H, O-H), 1647 (m, C=O), 1595 (m), 1493 (s), 1453 (m) (C=C, Ph), 1369 (m), 1201 (m), 774 (vs), 748 (vs), 690 (vs), 675 (sh) (C-H Aromatic out of plane). 1H NMR δ: 9.77 (s, 1H, CONH), 9.34 (s, 1H, OH), 8.51 (s, 1H, OH), 7.80 (d, 2H, 3J = 8.2, H11), 7.55 (s, 1H, PhNH), 7.44 (t, 2H, 3J = 7.1, H12), 7.41 (t, 1H, 3J = 7.1, H13), 7.18 (t, 2H, 3J = 7.6, H8), 6.95 (d, 1H, 3J = 7.6, H16), 6.90 (t, 2H, 3J = 7.6, H22), 6.83 (d, 2H, 3J = 7.6, H7), 6.80 (t, 1H, 3J = 7.6, H9), 6.57 (dd, 1H, 3J = 7.6, 4J = 2.9, H17), 6.54 (t, 1H, 3J = 7.6, H23), 6.51 (d, 1H, 3J = 2.9, H19), 6.13 (d, 2H, 3J = 7.6, H21), 5.55 (d, 1H, 3J = 12.3, H5), 4.95 (d, 1H, 3J = 12.3, H4). RMN 13C δ: 167.2 (CO), 149.7 (C15), 148.6 (C18), 147.4 (C3), 147.0 (C20), 145.5 (C6), 132.0 (C10), 128.9 (C13), 128.72 (C8), 128.70 (C12), 128.5 (C22), 125.9 (C11), 122.5 (C14), 119.8 (C9), 118.1 (C17), 115.7 (C19), 115.6 (C7), 115.3 (C23), 114.6 (C16), 111.9 (C21), 63.0 (C5), 54.2 (C4). Mass analysis [M-H]+ (m/z): 465.1932 found, 465.1927 calculated.
(±)-5-(5′-Bromo-2′-hydroxy-3-methoxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2h). Synthesized as described for 2a, from 0.300 g (0.83 mmol) of 1h, 0.35 mL of phenylhydrazine (3.6 mmol) to obtain 0.12 g (0.21 mmol, 26% yield, 99.29% purity) of a white fluorescent solid, mp. = 207–210 °C. IR (cm−1): 3500 (br), 3370, 3290 (br) (N-H, O-H), 1650 (m, C=O), 1600 (m), 1490 (s), 1440 (m) (C=C, Ph), 1420 (m, CH3), 1370 (m), 1270 (s), 750, 692 (vs, C-H Aromatic out of plane). 1H NMR δ: 9.88 (br, 1H, CONH), 9.70 (br, 1H, OH), 7.81 (d, 2H, 3J = 8.2, H11), 7.56 (s, 1H, PhNH), 7.44 (dd, 2H, 3J = 8.2, 4J = 7.6, H12), 7.42 (t, 1H, 3J = 7.6, H13), 7.20 (dd, 2H, 3J = 8.0, 3J = 7.6, H8), 7.08 (d, 1H, 4J = 2.3, H17), 6.93 (dd, 2H, 3J = 7.7, 3J = 7.0, H22), 6.92 (d, 2H, 3J = 8.0, H7), 6.83 (t, 1H, 3J = 7.6, H9), 6.74 (d, 1H, 4J = 2.3, H19), 6.60 (t, 1H, 3J = 7.0, H23), 6.13 (d, 2H, 3J = 7.7, H21), 5.62 (d, 1H, 3J = 12.0, H5), 5.00 (d, 1H, 3J = 12.0, H4), 3.90 (s, 3H, OCH3). 13C NMR δ: 167.3 (CO), 148.8 (C16), 148.75 (C15), 148.0 (C3), 145.6 (C20), 143.5 (C6), 132.1 (C10), 129.4 (C13), 129.3 (C12), 129.2 (C8), 128.8 (C22), 126.3 (C11), 124.9 (C14), 123.5 (C19), 120.5 (C9), 118.7 (C23), 114.8 (C7), 114.2 (C17), 112.1 (C21), 110.4 (C18), 62.7 (C5), 56.6 (OMe), 54.8 (C4). Mass analysis [M-H]+ (m/z): 557.1195 found, 557.1188 calculated.
(±)-5-(3′-Ethoxy-2′-hydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2i). Synthesized as described for 2a from 0.300 g (1.01 mmol) of 1i, 0.35 mL of phenylhydrazine (3.6 mmol) to obtain 0.16 g (0.32 mmol, 32% yield, 99.50% purity) of a white fluorescent solid, mp. = 205–207 °C. IR (cm−1): 3496 (w), 3320 (br), 3258, 3058 (br) (N-H, O-H), 1650 (m, C=O), 1600 (m), 1490 (m) 1470 (m) (C=C, Ph), 1440 (m, CH3), 1370 (m), 1270 (m), 750 (vs), 690 (vs), 650 (m) (C-H Aromatic out of plane). 1H NMR δ: 9.75 (s, 1H, CONH), 9.00 (s, 1H, OH), 7.81 (d, 2H, 3J = 8.2, H11), 7.55 (s, 1H, PhNH), 7.45 (dd, 2H, 3J = 8.2, 3J = 7.6, H12), 7.42 (t, 1H, 3J = 7.6, H13), 7.16 (dd, 2H, 3J = 8.2, 3J = 7.6, H8), 6.93 (d, 2H, 3J = 8.2, H7), 6.93 (d, 1H, 3J = 8.2, H17), 6.88 (dd, 2H, 3J = 8.2, 4J = 7.7, H22), 6.79 (t, 1H, 3J = 7.6, H9), 6.66 (d, 1H, 3J = 7.6, H19), 6.56 (t, 1H, 3J = 7.7, H23), 6.54 (dd, 1H, 3J = 7.6, 3J = 8.2, H18), 6.05 (d, 2H, 3J = 8.2, H21), 5.67 (d, 1H, 3J = 12, H5), 4.95 (d, 1H, 3J = 12, H4), 4.14 (q, 2H, 3J = 7.0, CH2), 1.45 (t, 3H, 3J = 7.0, CH3). RMN 13C δ: 167.6 (CO), 148.9 (C16), 147.6 (C15), 146.8 (C3), 145.8 (C20), 144.0 (C6), 132.4 (C10), 129.2 (C13), 129.1 (C12), 129.0 (C8), 128.7 (C22), 126.2 (C11), 122.9 (C14), 121.4 (C19), 120.2 (C9), 119.2 (C18), 118.4 (C23), 114.9 (C7), 112.2 (C21), 112.1 (C17), 64.4 (OCH2), 63.1 (C5), 54.7 (C4), 15.3 (CH3). Mass analysis [M-H]+ (m/z): 493.2244 found, 493.2240 calculated.

3.4. Modelling and In Silico Studies

3.4.1. Docking Simulations

Enantiomers (4S, 5S) and (4R, 5R) of DPCH derivatives, Figure S54, were drawn using CHEMSKETCH program 11.12; atomic connectivity was checked with GAUSS VIEW 3.0 and then geometrically optimized using Gaussian 09W at the AM1 level [93]. The catalytic domain-2 of HDAC6 (DD2-HDAC6) (PDB: 5G0J) was retrieved from previous work [94]. The 3D structure of DD2-HDAC6 was prepared using AutoDock Tools 1.5.6 [95]; polar hydrogen atoms and Kollman [96] charges were assigned for receptor and ligands. Validation of the method was performed with TSA with a root-mean-square deviation (RMSD) value of 2.05 Å, Figure S55. The grid box was centered on the receptor with grid points in the x, y, and z of 126 Å3, with a grid spacing of 0.375 Å3. A Lamarckian genetic algorithm was used as a scoring sample for a randomized population of 100 individuals, on which a 107 energy evaluations were done; 100 runs were performed. A focused molecular docking at Zn+2 coordinates was performed using AutoDock 4.2 and AutoDock4Zn force field, which has improved parameters to dock zinc proteins [97]. The most populated cluster conformations and the lowest free energy of binding values (ΔGb°) were selected as the most representative. Docking results of the DD2-HDAC6-ligand complexes were analyzed using AutoDock Tools 1.5.6 [98]. Figures were further processed using Pymol v.099 [99].

3.4.2. Theoretical ADME-Tox and Physicochemical Properties

The proposed compounds were submitted to determine their ADME-Tox properties on OSIRIS DataWarrior (v04.06.01) and Osiris Property Explorer [100]. The molecules were drawn using ChemBioDraw Ultra 12.0, and the simplified molecular input line entry specification (SMILES) codes for all compounds were obtained. The following properties were obtained from OSIRIS Property Explorer: mutagenic, tumorigenic, irritant, and reproductive effects—likewise, solubility in water (LogS) and topological surface area (TPSA) values. In the case of the Lipinski’s rules properties, these were determined from OSIRIS DataWarrior: molecular weight (MW), octanol–water partition coefficient (LogP), hydrogen acceptors (HA), hydrogen donors (HD), and rotatable bonds (RB). The size was measured by molar refractivity (MR) [68] and lipophilicity by the partition coefficient [101] parameters that were determined on ACD/ChemSketch and CS ChemDraw Pro v.6 software, respectively. All these biological, toxic, and physicochemical properties were compared with tubacin, TSA, and SAHA.

3.4.3. QSAR Analysis

QSAR was performed under QSAR-2D [81,102]. Estimation of the lipid solubility descriptor (π) values was performed by means of the following equation: π = log(PX/PH), where Px and PH are the partition coefficients of the substituted and leading compounds, respectively. The Hammett constant in the para position (σp) was utilized as the criterion of electronic effects [103]. Estimation of the steric descriptor (ES) values [104,105] was performed by means of the following equation: ES = log(MRX/MRH), where MRx and MRH are the molar refractivity values of the substituted and the leading compounds, respectively. The correlations were carried out through second-order polynomial regression analyses (y = Ax2 + Bx + C). The equation constants and parabolic correlation coefficient were analyzed under the Student’s test. The differences were considered significant for a minimal value of p < 0.05. Statistical tests were performed on Sigma Stat 3.5 software (Jandel Corp. SPSS INC. San Rafael, CA, USA).

3.5. In Vitro Assays

3.5.1. Cell Culture

The cancer cell lines used in this study were obtained from the American Type Tissue Culture Collection (ATCC), Rockville, MD, USA. MCF-7 and MDA-MB-231 are from BC cells, and 3T3/NIH and MCF10A cells were included as nonmalignant cells. BC cell lines and fibroblasts were grown in Dulbecco’s modified Eagle Medium (DMEM) high-glucose with phenol red. The culture medium was supplemented with 10% fetal bovine serum (FBS, BioWest, Miami, FL, USA) as well as 100 U/mL penicillin and 100 mg/mL streptomycin as antibiotic. MCF10A cells were cultured in DMEM/F-12 supplemented with 5% horse serum (Biowest, Miami, FL, USA), 20 ng/mL epidermal growth factor, 10 mg/mL insulin, and 500 ng/mL hydrocortisone. Cell cultures were incubated at 37 °C in a humidified atmosphere of 5% CO2 and 95% air. Cells were grown until 80% confluence, treated with trypsin-EDTA (1%) (4 mL, 5 min, 37 °C), and then collected with medium (4 mL). Cells were centrifuged (3 × 103 rpm, 10 min) and resuspended in medium (1–3 mL) and counted with CytoSmart cell counting (CytoSmart Technologies, Eindhoven, The Netherlands). Each cell line (10 × 103 cells per well) was cultured in 96-well plates and allowed to attach for 24 h before the assays. Then, cells were treated with the tested compounds at different concentrations (10–120 µM) for 48 h; all compounds were dissolved in DMSO to produce a final concentration of DMSO of (0.1%).

3.5.2. Cell Proliferation Assays

Cell proliferation was determined using the MTT [3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-tetrazolium bromide, Sigma] assay. For this purpose, MTT (0.500 mg mL−1), dissolved in phosphate buffered saline (PBS), was added to each well (after aspirating the medium) and incubated for 3 h at 37 °C in 5% CO2. The MTT/PBS was removed, and 100 µL of DMSO was added to each well to dissolve the formazan crystals. Absorbance was measured with a microplate reader (ThermoScientific, MultiskanTM Sky) at a wavelength of 550 nm. The quantity of formazan produced is directly proportional to the number of living cells. Results are expressed as the percentage of viable cells ± standard deviation in relation to the control (cell culture medium with 0.1% of DMSO), whose viability was designated as 100%. Each data point was performed in octuplicate in three independent experiments, and the results were reported as the mean absorption ± SD.

3.5.3. In Vitro HDAC6 Inhibition

The HDAC6 activity was measured using the Fluor de Lys-HDAC6 assay kit (ENZO Life Sciences). The method consists of deacetylation of the substrate (Fluor de lys-SIRT1) in the presence of human recombinant HDAC6. Then, the deacetylated substrate is incubated at room temperature for 45 min with Fluor de Lys-developer II to generate a fluorophore that can be measured by fluorescence (Fluorescence Spectrometer LS 55 PerkinElmer) at an excitation/emission wavelength of 360/460 nm. The HDAC6 inhibition by TSA was determined using different concentrations (0.05, 5, 50, and 250 nM), while the inhibition by compound 2b was evaluated with five concentrations (0.5, 5, 10, 35, and 50 µM). The HADC6 activity was expressed in percentage, and it was calculated with the following equation:
%HDAC6 Activity = absorbance of inhibition × 100/ absorbance of control

3.5.4. Confocal Fluorescence Microscopy

An aliquot of MCF-7 breast cancer cells was seeded in petri dishes with a coverslip in clear media (supplemented media phenol-red-free). The cells were incubated at 37 °C overnight in 5% CO2. Once the cells were adhered, the tested compound was added at 10 µM concentrations for 30 min. Then, the cells were washed several times with PBS and immersed in cold ethanol. Micrographs were acquired with confocal laser scanning microscope LSM 710 NLO, Carl Zeiss, Germany.

3.5.5. Wound Closure Assay

BC cells (1.5 × 105) cells were 24-well plated and allowed to reach 100% confluence. Cell monolayers were scratched with a 200 µL sterile pipette tip to form wound gaps, and the media and cell debris were carefully aspirated. Culture media was replaced and compounds 2b and 2c at 15 µM were added. The wound closure was monitored by microscopy at 16, 24, and 48 h. The wound area was measured by quadruplicate in two independent experiments and expressed as percentage of the control (cells culture medium with 0.1% of DMSO).

3.5.6. DPPH Assay (2,2-Diphenyl-1-picrylhydrazyl)

Into a 96-well plate, we poured 100 μL of DPPH 0.20 mM in absolute methanol and 100 μL of the appropriate compound (6.25, 12.5, 25, 50, 100, 200 µM final concentrations) dissolved in DMSO, the mixtures were incubated for 30 min at room temperature protected from light [106]. Each assay was performed in triplicate with ascorbic acid (AA) as a standard. The absorbance was measured at 517 nm in a transparent 96-well test microplate (Multiskan-EX Thermo Scientific, Waltham, MA, USA). The results are shown as percentage of DPPH radical reduced at each concentration. Therefore, the antioxidant activity (DPPH scavenging) of each compound was calculated by the following equation: [1 − (A1 − A2)/(ADPPH − AS)] × 100, where: A1 = absorbance of the compound with DPPH, A2 = absorbance of the compound, ADPPH = absorbance of DPPH (diluted 1:1 with solvent) and AS = absorbance of the solvent. The experiments were performed in triplicate with several concentrations and the IC50 values were calculated using GraphPad Prism 8.

3.6. Statistical Analysis

Where needed, results were compared by one-way ANOVA with Dunnett post-test. GraphPad Prism version 8 for Windows was used for statistical analysis. A difference was considered statistically significant if p ≤ 0.05. The half-maximal inhibitory concentration (IC50) was calculated from the dose–response curves through a logarithmic analysis of HillSlope.

4. Conclusions

In summary, the synthesis and chemical characterization of nine new 4,5-dihydropyazole-carbohydrazide derivatives with dual antioxidant and antiproliferative activities on BC cell lines are described. The synthesized compounds had more favorable physicochemical and ADME-Tox characteristics than tubacin, but were comparable to TSA and SAHA, the known HDAC6 inhibitors. An antiproliferative effect against cancer cell lines MCF-7 and MDA-MB-231, as well as low cytotoxicity against normal breast cells, was demonstrated. In particular, compounds with R = H (2a), 6-OMe (2c), and 8-OEt (2i) showed the smallest IC50 values against BC cells and the smallest cytotoxicity towards nonmalignant breast cells, being capable of crossing the cell membrane. Furthermore, compounds 2b (6-Cl) and 2c (6-OMe) diminished the motility of TNBC cells and inhibited the human HDAC6 with free binding energies like TSA and SAHA. QSAR supported a size effect, probably by blocking the entrance of the DD2 catalytic domain, with close similarity to the mode of action of tubacin. Finally, these compounds are effective dual anticancer–antioxidant agents with reduced cytotoxicity in healthy cells. Further studies on other HDACs isoforms are currently in progress.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ph15060690/s1, Figure S1: NOE spectra of compound 5-(2-hydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2a); Figure S2: Binding conformation around DD2-HDAC6 catalytic domain obtained through blind docking; Figures S3–S4: Half-maximal inhibitory concentration 50 (IC50) in BC cell lines MCF-7 and MDA-MB-231; Figures S5–S8: Antiproliferative activity in the non-malignant cell lines 3T3/NIH and MCF10A; Figure S9: (a) π values and σH values of DPCH derivatives on healthy cellular line 3T3/NIH; Figure S10: Comparison of the radical-scavenging activity of compounds 2ai and ascorbic acid; Figures S11–S28: 1H and 13C NMR spectra of compound 2ai in DMSO-d6; Figures S29–S30: COSY and HETCOR spectra of compound 5-(2-hydroxyphenyl)-N′,1,3-triphenyl-4,5-dihydro-1H-pyrazole-4-carbohydrazide (2a); Figures S31–S39: IR spectra of compounds 2ai; Figures S40–S44: Mass spectra of compounds 2ai; Figures S45–S53: HPLC chromatograms-purity of compounds 2ai; Figure S54: Enantiomers (4S, 5S) A and (4R, 5R) B of modelled 4,5-dihydropyrazole derivatives; Figure S55: Overlay of TSA in the DD2-HDAC6 domain with an RMSD value of 2.05; Table S1: Bond lengths (Å), Bond and torsion angles (°) of 2a; Table S2: Hydrogen bonding geometry parameters of 2a; Table S3: Free binding energy ΔGb° (kcal/moL) and Kd (µM) values obtained by docking the DD2-HDAC6 domain with 4,5-dihydropyrazole derivatives 2ai; Table S4: Interactions among of the 4,5-dihydropyrazole derivatives compared with tubacin, TSA, and SAHA with the DD2-HDAC6 structure (PDB: 5G0J); Table S5: Toxicity profile of the 4,5-dihydropyrazole derivatives compared with tubacin and TSA; Table S6: Crystal data and details of the structure determination for 2a.

Author Contributions

Conceptualization, J.C.-B., M.C.R.-H. and I.I.P.-M.; data curation, I.I.P.-M.; formal analysis, I.B.-R., A.M.R.-A., Y.S.-L., J.E.M.-W., F.J.M.-M., I.M.O.-C., J.R.G.-S., J.A.G.-S. and M.B.; funding acquisition, J.C.-B., M.C.R.-H. and I.I.P.-M.; methodology, I.B.-R., A.M.R.-A., Y.S.-L., J.E.M.-W., F.J.M.-M., I.M.O.-C., J.R.G.-S. and M.B.; resources, I.I.P.-M.; writing—original draft, I.B.-R.; writing—review and editing, J.C.-B., M.C.R.-H. and I.I.P.-M. All authors have read and agreed to the published version of the manuscript.

Funding

Partially supported by Consejo Nacional de Ciencia y Tecnología (CONACYT, grants 255354 and 254600) and Secretaría de Investigación y Posgrado del Instituto Politécnico Nacional (SIP, grants 20211070 and 20201274). Authors thank Dr. Susana Rojas-Lima (UAEH) for access to the X-ray diffractometer.

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 that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. The authors declare no conflict of interest.

References

  1. World Health Organization (WHO), Breast Cancer. 2021. Available online: https://www.who.int/news-room/fact-sheets/detail/breast-cancer (accessed on 24 January 2022).
  2. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global Cancer Statistics: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Claude-Taupin, A.; Boyer-Guittaut, M.; Delage-Mourroux, R.; Hervouet, E. Use of Epigenetic Modulators as a Powerful Adjuvant for Breast Cancer Therapies. In Cancer Epigenetics; Methods in Molecular Biology; Humana Press: New York, NY, USA, 2015; Volume 1238, pp. 487–509. [Google Scholar] [CrossRef]
  4. Perou, C.M.; Sorlie, T.; Eisen, M.B.; van de Rijn, M.; Jeffrey, S.S.; Rees, C.A.; Pollack, J.R.; Ross, D.T.; Johnsen, H.; Akslen, L.A.; et al. Molecular portraits of human breast tumours. Nature 2000, 406, 747–752. [Google Scholar] [CrossRef] [PubMed]
  5. Buocikova, V.; Rios-Mondragon, I.; Pilalis, E.; Chatziioannou, A.; Miklikova, S.; Mego, M.; Pajuste, K.; Rucins, M.; El-Yamani, N.; Longhin, E.M.; et al. Epigenetics in Breast Cancer Therapy—New Strategies and Future Nanomedicine Perspectives. Cancers 2020, 12, 3622. [Google Scholar] [CrossRef] [PubMed]
  6. Marks, D.L.; Olson, R.L.; Fernandez-Zapico, M.E. Epigenetic control of the tumor microenvironment. Epigenomics 2016, 8, 1671–1687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Giacinti, L.; Claudio, P.P.; Lopez, M.; Giordano, A. Epigenetic information and estrogen receptor alpha expression in breast cancer. Oncologist 2006, 11, 1–8. [Google Scholar] [CrossRef] [Green Version]
  8. Sadoul, K.; Boyault, C.; Pabion, M.; Khochbin, S. Regulation of protein turnover by acetyltransferases and deacetylases. Biochimie 2008, 90, 306–312. [Google Scholar] [CrossRef]
  9. Dekker, F.J.; Haisma, H.J. Histone acetyl transferases as emerging drug targets. Drug Discov. Today 2009, 14, 942–948. [Google Scholar] [CrossRef]
  10. Grozinger, C.M.; Hassig, C.A.; Schreiber, S.L. Three proteins define a class of human histone deacetylases related to yeast Hda1p. Proc. Natl. Acad. Sci. USA 1999, 96, 4868–4873. [Google Scholar] [CrossRef] [Green Version]
  11. Marks, P.; Rifkind, R.A.; Richon, V.M.; Breslow, R.; Miller, T.; Kelly, W.K. Histone deacetylases and cancer: Causes and therapies. Nat. Rev. Cancer 2001, 1, 194–202. [Google Scholar] [CrossRef]
  12. Abel, T.; Zukin, R.S. Epigenetic targets of HDAC inhibition in neurodegenerative and psychiatric disorders. Curr. Opin. Pharmacol. 2008, 8, 57–64. [Google Scholar] [CrossRef] [Green Version]
  13. D’Mello, S.R. Histone deacetylases as targets for the treatment of human neurodegenerative diseases. Drug News Perspect. 2009, 22, 513–524. [Google Scholar] [CrossRef]
  14. Mottamal, M.; Zheng, S.; Huang, T.L.; Wang, G. Histone Deacetylase Inhibitors in Clinical Studies as Templates for New Anticancer Agents. Molecules 2015, 20, 3898–3941. [Google Scholar] [CrossRef] [Green Version]
  15. Bolden, J.E.; Peart, M.J.; Johnstone, R.W. Anticancer activities of histone deacetylase inhibitors. Nat. Rev. Drug Discov. 2006, 5, 769–784. [Google Scholar] [CrossRef]
  16. Shakespear, M.R.; Halili, M.A.; Irvine, K.M.; Fairlie, D.P.; Sweet, M.J. Histone deacetylases as regulators of inflammation and immunity. Trends Immunol. 2011, 32, 335–343. [Google Scholar] [CrossRef]
  17. Chuang, D.M.; Leng, Y.; Marinova, Z.; Kim, H.J.; Chiu, C.T. Multiple roles of HDAC inhibition in neurodegenerative conditions. Trends Neurosci. 2009, 32, 591–601. [Google Scholar] [CrossRef] [Green Version]
  18. Park, J.-H.; Jung, Y.; Kim, T.Y.; Kim, S.G.; Jong, H.-S.; Lee, J.W.; Kim, D.-K.; Lee, J.-S.; Kim, N.K.; Kim, T.-Y.; et al. Class I histone deacetylase-selective novel synthetic inhibitors potently inhibit human tumor proliferation. Clin. Cancer Res. 2004, 10, 5271–5281. [Google Scholar] [CrossRef] [Green Version]
  19. Zhang, Y.; Li, N.; Caron, C.; Matthias, G.; Hess, D.; Khochbin, S.; Matthias, P. HDAC-6 interacts with and deacetylates tubulin and microtubules in vivo. EMBO J. 2003, 22, 1168–1179. [Google Scholar] [CrossRef] [Green Version]
  20. Grozinger, C.M.; Schreiber, S.L. Deacetylase enzymes: Biological functions and the use of small-molecule inhibitors. Chem. Biol. 2002, 9, 3–16. [Google Scholar] [CrossRef] [Green Version]
  21. Carey, N.; La Thangue, N.B. Histone deacetylase inhibitors: Gathering pace. Curr. Opin. Pharmacol. 2006, 6, 369–375. [Google Scholar] [CrossRef]
  22. Bertrand, P. Inside HDAC with HDAC inhibitors. Eur. J. Med. Chem. 2010, 45, 2095–2116. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Gilquin, B.; Khochbin, S.; Matthias, P. Two catalytic domains are required for protein deacetylation. J. Biol. Chem. 2006, 281, 2401–2404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhang, X.; Yuan, Z.; Zhang, Y.; Yong, S.; Salas-Burgos, A.; Koomen, J.; Olashaw, N.; Parsons, J.T.; Yang, X.-J.; Dent, S.R.; et al. HDAC6 modulates cell motility by altering the acetylation level of cortactin. Mol. Cell 2007, 27, 197–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Aoyagi, S.; Archer, T.K. Modulating molecular chaperone Hsp90 functions through reversible acetylation. Trends Cell Biol. 2005, 15, 565–567. [Google Scholar] [CrossRef] [PubMed]
  26. Deakin, N.O.; Turner, C.E. Paxillin inhibits HDAC6 to regulate microtubule acetylation, Golgi structure and polarized migration. J. Cell Biol. 2014, 206, 395–413. [Google Scholar] [CrossRef] [Green Version]
  27. Kovacs, J.J.; Murphy, P.J.; Gaillard, S.; Zhao, X.; Wu, J.T.; Nicchita, C.V.; Yoshida, M.; Toft, D.O.; Pratt, W.B.; Yao, T.P. HDAC6 regulates Hsp90 acetylation and chaperone-dependent activation of glucocorticoid receptor. Mol. Cell 2005, 18, 601–607. [Google Scholar] [CrossRef]
  28. Zhang, Z.; Yamashita, H.; Toyama, T.; Sugiura, H.; Omoto, Y.; Ando, Y.; Mita, K.; Hamaguchi, M.; Hayashi, S.-I.; Iwase, H. HDAC6 expression is correlated with better survival in breast cancer. Clin. Cancer Res. 2004, 10, 6962–6968. [Google Scholar] [CrossRef] [Green Version]
  29. Seidel, C.; Schnekenburger, M.; Dicato, M.; Diederich, M. Histone deacetylase 6 in health and disease. Epigenomics 2015, 7, 103–118. [Google Scholar] [CrossRef] [Green Version]
  30. Roche, J.; Bertrand, P. Inside HDACs with more selective HDAC inhibitors. Eur. J. Med. Chem. 2016, 121, 451–483. [Google Scholar] [CrossRef]
  31. Zou, H.; Wu, Y.; Navre, M.; Sang, B.-C. Characterization of the two catalytic domains in histone deacetylase 6. Biochem. Biophys. Res. Commun. 2006, 341, 45–50. [Google Scholar] [CrossRef]
  32. Asthana, J.; Kapoor, S.; Mohan, R.; Panda, D. Inhibition of HDAC6 deacetylase activity increases its binding with microtubules and suppresses microtubule Dynamic instability in MCF-7 cells. J. Biol. Chem. 2013, 288, 22516–22526. [Google Scholar] [CrossRef] [Green Version]
  33. Matsuyama, A.; Shimazu, T.; Sumida, Y.; Saito, A.; Yoshimatsu, Y.; Seigneurin-Berny, D.; Osada, H.; Komatsu, Y.; Nishino, N.; Khochbin, S.; et al. In vivo destabilization of Dynamic microtubules by HDAC6 mediated deacetylation. EMBO J. 2002, 21, 6820–6831. [Google Scholar] [CrossRef] [Green Version]
  34. Dallavalle, S.; Pisano, C.; Zunino, F. Development and therapeutic impact of HDAC6-selective inhibitors. Biochem. Pharmacol. 2012, 84, 756–765. [Google Scholar] [CrossRef]
  35. Haggarty, S.J.; Koeller, K.M.; Wong, J.V.; Grozinger, C.M.; Schreiber, S.L. Domain-selective small-molecule inhibitor of histone deacetylase 6 (HDAC6)-mediated tubulin deacetylation. Proc. Natl. Acad. Sci. USA 2003, 100, 4389–4394. [Google Scholar] [CrossRef] [Green Version]
  36. Batchu, S.N.; Brijmohan, A.S.; Advani, A. The therapeutic hope for HDAC6 inhibitors in malignancy and chronic disease. Clin. Sci. 2016, 130, 987–1003. [Google Scholar] [CrossRef]
  37. Halliwell, B.; Gutteridge, J.M.C. Chapter 5 Oxidative Stress and Redox Regulation: Adaptation, Damage, Repair, Senescence and Death. In Free Radicals in Biology and Medicine, 5th ed.; Oxford University Press: Oxford, UK, 2015. [Google Scholar] [CrossRef]
  38. Ambrosone, C.B. Oxidants and Antioxidants in Breast Cancer. Antioxid. Redox Signal. 2000, 2, 903–917. [Google Scholar] [CrossRef]
  39. Shirwaikar, A.; Shirwaikar, A.; Rajendran, K.; Punitha, I.S.R. In Vitro Antioxidant Studies on the Benzyl Tetra Isoquinoline Alkaloid Berberine. Biol. Pharm. Bull. 2006, 29, 1906–1910. [Google Scholar] [CrossRef] [Green Version]
  40. Fleischauer, A.T.; Simonsen, N.; Arab, L. Antioxidant Supplements and Risk of Breast Cancer Recurrence and Breast Cancer-Related Mortality Among Postmenopausal Women. Nutr. Cancer 2003, 46, 15–22. [Google Scholar] [CrossRef]
  41. Greenlee, H.; Gammon, M.D.; Abrahamson, P.E.; Gaudet, M.M.; Terry, M.B.; Hershman, D.L.; Desai, M.; Teitelbaum, S.L.; Neugut, A.I.; Jacobson, J.S. Prevalence and Predictors of Antioxidant Supplement Use During Breast Cancer Treatment. Cancer 2009, 115, 3271–3282. [Google Scholar] [CrossRef] [Green Version]
  42. Fernandez-Lazaro, C.I.; Martínez-González, M.A.; Aguilera-Buenosvinos, I.; Gea, A.; Ruiz-Canela, M.; Romanos-Nanclares, A.; Toledo, E. Dietary Antioxidant Vitamins and Minerals and Breast Cancer Risk: Prospective Results from the SUN Cohort. Antioxidants 2021, 10, 340. [Google Scholar] [CrossRef]
  43. Goh, J.; Pettan-Brewer, C.; Enns, L.; Fatemie, S.; Ladiges, W. Are Exercise and Mitochondrial Antioxidants Compatible in the Treatment of Invasive Breast Cancer? Bioenergy Open Access 2012, 1, 101. [Google Scholar] [CrossRef] [Green Version]
  44. Kumar, M.R.; Dhayabaran, V.V.; Sudhapriya, N.; Manikandan, A.; Gideon, D.A.; Annapoorani, S. p-TSA.H2O mediated one-pot, multi-component synthesis of isatin derived imidazoles as dual-purpose drugs against inflammation and cancer. Bioorg. Chem. 2020, 102, 104046. [Google Scholar] [CrossRef]
  45. Harisha, S.; Keshavayya, J.; Prasanna, S.M.; Hoskeri, H.J. Synthesis, characterization, pharmacological evaluation and molecular docking studies of benzothiazole azo derivatives. J. Mol. Struct. 2020, 1218, 128477. [Google Scholar] [CrossRef]
  46. Chatterjee, A.; Ronghe, A.; Padhye, S.B.; Spade, D.A.; Bhat, N.K.; Bhat, H.K. Antioxidant activities of novel resveratrol analogs in breast cancer. J. Biochem. Mol. Toxicol. 2018, 32, e21925. [Google Scholar] [CrossRef] [PubMed]
  47. Sheu, M.-T.; Jhan, H.-J.; Hsieh, C.-M.; Wang, C.-J.; Ho, H.-O. Efficacy of Antioxidants as a Complementary and Alternative Medicine (CAM) in Combination with the Chemotherapeutic Agent Doxorubicin. Integr. Cancer Ther. 2015, 14, 184–195. [Google Scholar] [CrossRef] [PubMed]
  48. Chimenti, F.; Bolasco, A.; Manna, F.; Secci, D.; Chimenti, P.; Befani, O.; Turin, P.; Giovannini, V.; Mondovi, B.; Cirilli, R.; et al. Synthesis and selective inhibitory activity of 1-acetyl-3,5-diphenyl-4,5-dihydro-(1H)-pyrazole derivatives against monoamine oxidase. J. Med. Chem. 2004, 47, 2071–2074. [Google Scholar] [CrossRef]
  49. Girisha, K.S.; Kalluraya, B.; Narayana, V. Padmashree, Synthesis and pharmacological study of 1-acetyl/propil-3-aryl-5-(5-chloro-3-methyl-1-phenyl-1H-pyrazol-4-yl)-2-pyrazoline. Eur. J. Med. Chem. 2010, 45, 4640–4644. [Google Scholar] [CrossRef]
  50. Bao, H.; Zhang, Q.; Zhu, Z.; Xu, H.; Ding, F.; Wang, M.; Du, Y.; Yan, Z. BHX, a novel pyrazoline derivative, inhibits breast cancer cell invasion by reversing the epithelial mesenchymal transition and down-regulating Wnt/β-catenin signalling. Sci. Rep. 2017, 7, 9153. [Google Scholar] [CrossRef]
  51. Yan, Z.; Zhu, Z.; Wang, J.; Sun, J.; Chen, Y.; Yang, G.; Chen, W.; Deng, Y. Synthesis, characterization, and evaluation of a novel inhibitor of WNT/β-catenin signaling pathway. Mol. Cancer 2013, 12, 116. [Google Scholar] [CrossRef] [Green Version]
  52. Park, S.Y.; Jun, J.A.; Jeong, K.J.; Heo, H.J.; Sohn, J.S.; Lee, H.Y.; Park, C.G.; Kang, J. Histone deacetylases 1, 6 and 8 are critical for invasion in breast cancer. Oncol. Rep. 2011, 25, 1677–1681. [Google Scholar] [CrossRef] [Green Version]
  53. Butler, K.V.; Kalin, J.; Brochier, C.; Vistoli, G.; Langley, B.; Kozikowski, A.P. Rational Design and Simple Chemistry Yield a Superior, Neuroprotective HDAC6 Inhibitor, Tubastatin A. J. Am. Chem. Soc. 2010, 132, 10842–10846. [Google Scholar] [CrossRef] [Green Version]
  54. Osko, J.D.; Christianson, D.W. Structural determinants of affinity and selectivity in the binding of inhibitors to histone deacetylase 6. Bioorg. Med. Chem. Lett. 2020, 30, 127023. [Google Scholar] [CrossRef]
  55. Sixto-López, Y.; Gómez-Vidal, J.A.; de Pedro, N.; Bello, M.; Rosales-Hernández, M.C.; Correa-Basurto, J. Hydroxamic acid derivatives as HDAC1, HDAC6 and HDAC8 inhibitors with antiproliferative activity in cancer cell lines. Sci. Rep. 2020, 10, 10462. [Google Scholar] [CrossRef]
  56. Estiu, G.; Greenberg, E.; Harrison, C.B.; Kwiatkowski, N.P.; Mazitschek, R.; Bradner, J.E.; Wiest, O. Structural Origin of Selectivity in Class II-Selective Histone Deacetylase Inhibitors. J. Med. Chem. 2008, 51, 2898–2906. [Google Scholar] [CrossRef]
  57. Kerr, J.S.; Galloway, S.; Lagrutta, A.; Armstrong, M.; Miller, T.; Richon, V.M.; Andrews, P.A. Nonclinical safety assessment of the histone deacetylase inhibitor vorinostat. Int. J. Toxicol. 2010, 29, 3–19. [Google Scholar] [CrossRef]
  58. Shen, S.; Kozikowski, A.P. Why Hydroxamates May Not Be the Best Histone Deacetylase Inhibitors—What Some May Have Forgotten or Would Rather Forget? ChemMedChem 2016, 11, 15–21. [Google Scholar] [CrossRef] [Green Version]
  59. Padilla-Martínez, I.I.; Flores-Larios, I.Y.; García-Báez, E.V.; González, J.; Cruz, A.; Martínez-Martínez, F.J. X-ray Supramolecular structure, NMR spectroscopy and synthesis of 3-methyl-1-phenyl-1H-chromeno [4,3-c]pyrazol-4-ones formed by the unexpected cyclization of 3-[1-(phenyl-hydrazono)ethyl]–chromen-2-ones. Molecules 2011, 16, 915–932. [Google Scholar] [CrossRef] [Green Version]
  60. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2016, B72, 171–179. [Google Scholar] [CrossRef]
  61. Liu, L.; Ji, Y.-L.; Jia, D.-Z.; Yu, K.-B. Synthesis and crystal structure of supramolecular compound of 4-(a’ hydroxybenzoylhydrazinyl)benzal/ethylidene-5-methyl-2-phenyl-2,4-dihydropyrazol-3-one. Huaxue Xuebao 2003, 61, 893–900. [Google Scholar] [CrossRef]
  62. Li, S.S.; Zhao, C.; Zhang, G.; Xu, Q.; Liu, Q.; Zhao, W.; Chou, C.J.; Zhang, Y. Development of selective HDAC6 inhibitors with in vitro and in vivo anti-multiple myeloma activity. Bioorg. Chem. 2021, 116, 105278. [Google Scholar] [CrossRef]
  63. Sixto-López, Y.; Bello, M.; Rodríguez-Fonseca, R.A.; Rosales-Hernández, M.C.; Martínez-Archundia, M.; Gómez-Vidal, J.A.; Correa-Basurto, J. Searching the conformational complexity and binding properties of HDAC6 through docking and molecular dynamic simulations. J. Biomol. Struct. Dyn. 2017, 35, 2794–2814. [Google Scholar] [CrossRef]
  64. Kaliszczak, M.; Trousil, S.; Åberg, O.; Perumal, M.; Nguyen, Q.-D.; Aboagye, E.O. A novel small molecule hydroxamate preferentially inhibits HDAC6 activity and tumour growth. Br. J. Cancer 2013, 108, 342–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Lernoux, M.; Schnekenburge, M.; Dicato, M.; Diederich, M. Anti-cancer effects of naturally derived compounds targeting histone deacetylase 6-related pathways. Pharmacol. Res. 2017, 129, 337–356. [Google Scholar] [CrossRef] [PubMed]
  66. Veber, D.F.; Johnson, S.R.; Cheng, H.Y.; Smith, B.R.; Ward, K.W.; Kopple, K.D. Molecular properties that influence the oral bioavailability of drug candidates. J. Med. Chem. 2002, 45, 2615–2623. [Google Scholar] [CrossRef]
  67. Kubinyi, H. Strategies and recent technologies in drug discovery. Die Pharm. 1995, 50, 647–662. [Google Scholar]
  68. Ghose, A.K.; Crippen, G.M. Atomic physicochemical parameters for three-dimensional-structure-directed quantitative structure-activity relationships. 2. Modeling dispersive and hydrophobic interactions. J. Chem. Inf. Comput. Sci. 1987, 27, 21–35. [Google Scholar] [CrossRef] [PubMed]
  69. Zhao, Y.H.; Abraham, M.H.; Le, J.; Hersey, A.; Luscombe, C.N.; Beck, G.; Sherbone, B.; Cooper, I. Rate-Limited Steps of Human Oral Absorption and QSAR Studies. Pharm. Res. 2002, 19, 1446–1457. [Google Scholar] [CrossRef]
  70. Di Micco, S.; Chini, M.G.; Terracciano, S.; Bruno, I.; Riccio, R.; Bifulco, G. Structural basis for the design and synthesis of selective HDAC inhibitors. Bioorg. Med. Chem. 2013, 21, 3795–3807. [Google Scholar] [CrossRef] [PubMed]
  71. Hughes, D.; Andersson, D.I. Evolutionary consequences of drug resistance: Shared principles across diverse targets and organisms. Nat. Rev. Genet. 2015, 16, 459–471. [Google Scholar] [CrossRef]
  72. Yan, C.; Li, T.S. Dual role of mitophagy in cancer drug resistance. Anticancer Res. 2018, 38, 617–621. [Google Scholar] [CrossRef] [Green Version]
  73. Santiago-O’Farrill, J.M.; Weroha, S.J.; Hou, X.; Oberg, A.L.; Heinzen, E.P.; Maurer, M.J.; Pang, L.; Rask, P.; Amaravadi, R.K.; Becker, S.E.; et al. Poly(adenosine diphosphate ribose) polymerase inhibitors induce autophagy-mediated drug resistance in ovarian cancer cells, xenografts, and patient-derived xenograft models. Cancer 2019, 126, 894–907. [Google Scholar] [CrossRef]
  74. Wu, B.; Fathi, S.; Mortley, S.; Mohiuddin, M.; Jang, Y.C.; Oyelere, A.K. Pyrimethamine conjugated histone deacetylase inhibitors: Design, synthesis and evidence for triple negative breast cancer selective cytotoxicity. Bioorg. Med. Chem. 2020, 28, 115345. [Google Scholar] [CrossRef]
  75. Liang, X.; Tang, S.; Liu, X.; Liu, Y.; Xu, Q.; Wang, X.; Saidahmatov, A.; Li, C.; Wang, J.; Zhou, Y.; et al. Discovery of Novel Pyrrolo [2,3-d]pyrimidine-based Derivatives as Potent JAK/HDAC Dual Inhibitors for the Treatment of Refractory Solid Tumors. J. Med. Chem. 2021, 65, 1243–1264. [Google Scholar] [CrossRef]
  76. Bradner, J.E.; West, N.; Grachan, M.L.; Greenberg, E.F.; Haggarty, S.J.; Warnow, T.; Mazitschek, R. Chemical phylogenetics of histone deacetylases. Nat. Chem. Biol. 2010, 6, 238–243. [Google Scholar] [CrossRef]
  77. Zhao, C.; Gao, J.; Zhang, L.; Su, L.; Luan, Y. Novel HDAC6 selective inhibitors with 4-aminopiperidine-1-carboxamide as the core structure enhanced growth inhibitory activity of bortezomib in MCF-7 cells. Biosci. Trends 2019, 13, 91–97. [Google Scholar] [CrossRef] [Green Version]
  78. Wang, X.X.; Wan, R.-Z.; Liu, Z.-P. Recent advances in the discovery of potent and selective HDAC6 inhibitors. Eur. J. Med. Chem. 2018, 143, 1406–1418. [Google Scholar] [CrossRef]
  79. Taft, R.W., Jr. Polar and steric substituent constants for aliphatic and o-benzoate groups from rates of sterification and hydrolysis of esters. J. Am. Chem. Soc. 1952, 74, 3120–3128. [Google Scholar] [CrossRef]
  80. Hansch, C.; Fujita, T. p-σ-π Analysis. A method for the correlation of biological activity and chemical structure. J. Am. Chem. Soc. 1963, 86, 1616–1626. [Google Scholar] [CrossRef]
  81. Hammett, L.P. The effect of structure upon the reactions of organic compounds. Benzene derivatives. J. Am. Chem. Soc. 1937, 59, 96–103. [Google Scholar] [CrossRef]
  82. Rivera-Antonio, A.; Rosales-Hernández, M.C.; Balbuena-Rebolledo, I.; Santiago-Quintana, J.M.; Mendieta-Wejebe, J.E.; Correa-Basurto, J.; García-Vázquez, J.B.; García-Báez, E.V.; Padilla-Martínez, I.I. Myeloperoxidase Inhibitory and Antioxidant Activities of (E)-2- Hydroxy-aminocinnamic Acids Obtained through Microwave-Assisted Synthesis. Pharmaceuticals 2021, 14, 513. [Google Scholar] [CrossRef]
  83. Roy, D.; Chakraborty, A.; Ghosh, R. Perimidine based selective colorimetric and fluorescent turn-off chemosensor of aqueous Cu2+: Studies on its antioxidant property along with its interaction with calf thymus-DNA. RSC Adv. 2017, 7, 40563–40570. [Google Scholar] [CrossRef] [Green Version]
  84. Losada-Echeberría, M.; Herranz-López, M.; Micol, V.; Barrajón-Catalán, E. Polyphenols as Promising Drugs against Main Breast Cancer Signatures. Antioxidants 2017, 6, 88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. CrysAlis RED; Oxford Diffraction Ltd.: Abingdon, UK, 2006.
  86. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  87. Farrugia, L.J. WinGX and ORTEP for Windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  88. Spek, A.L. PLATON S-QUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 9–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Macrae, C.F.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Shields, G.P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury: Visualization and Analysis of Crystal Structures. J. Appl. Crystallogr. 2006, 39, 453–457. [Google Scholar] [CrossRef] [Green Version]
  90. Matos, M.J.; Vazquez-Rodriguez, S.; Uriarte, E.; Santana, L.; Viña, D. MAO inhibitory activity modulation: 3-Phenylcoumarins versus 3-benzoylcoumarins. Bioorg. Med. Chem. Lett. 2011, 21, 4224–4227. [Google Scholar] [CrossRef] [PubMed]
  91. Rao, H.S.P.; Sivakumar, S. Condensation of α-Aroylketene Dithioacetals and 2-Hydroxyarylaldehydes Results in Facile Synthesis of a Combinatorial Library of 3-Aroylcoumarins. J. Org. Chem. 2006, 71, 8715–8723. [Google Scholar] [CrossRef] [PubMed]
  92. Perez-Cruz, F.; Vazquez-Rodriguez, S.; Joao-Matos, M.; Herrera-Morales, A.; Villamena, F.A.; Das, A.; Gopalakrishnan, B.; Olea-Azar, C.; Santana, L.; Uriarte, E. Synthesis and Electrochemical and Biological Studies of Novel Coumarin−Chalcone Hybrid Compounds. J. Med. Chem. 2013, 56, 6136–6145. [Google Scholar] [CrossRef]
  93. Frisch, M.J.T.G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennuci, B.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian v09.01; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  94. Sixto-López, Y.; Bello, M.; Correa-Basurto, J. Structural and energetic basis for the inhibitory selectivity of both catalytic domains of dimeric HDAC6. J. Biomol. Struct. Dyn. 2019, 37, 4701–4720. [Google Scholar] [CrossRef]
  95. Gasteiger, J.; Marsili, M. Iterative partial equalization of orbital electronegativity—A rapid access to atomic charges. Tetrahedron 1980, 36, 3219–3228. [Google Scholar] [CrossRef]
  96. Singh, U.C.; Kollman, P.A. An approach to computing electrostatic charges for molecules. J. Comput. Chem. 1984, 5, 129–145. [Google Scholar] [CrossRef]
  97. Santos-Martins, D.; Forli, S.; Ramos, M.J.; Olson, A.J. AutoDock4(Zn): An improved AutoDock force field for small-molecule docking to zinc metalloproteins. J. Chem. Inf. Model. 2014, 54, 2371–2379. [Google Scholar] [CrossRef] [Green Version]
  98. Morris, G.M.; Huey, R.; Lindstrom, W.; Sanner, M.F.; Belew, R.K.; Goodsell, D.S.; Olson, A.J. AutoDock4 and AutoDockTools4: Automated docking with selective receptor flexibility. J. Comput. Chem. 2009, 30, 2785–2791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. DeLano, W.L.; Lam, J.W. The PyMOL Molecular Graphics System; Schrodinger, L., Ed.; DeLano Scientific: San Francisco, CA, USA, 2002; Available online: http://www.pymol.org/funding.html (accessed on 24 January 2022).
  100. 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]
  101. Morán, J.R.; Jiménez, H.A.; Gómez, R.; Arellano, M.G.; Quintana, D.; Guevara, J.A. Correlation study of antibacterial activity and spectrum of Penicillins through a structure-activity relationship analysis. Med. Chem. Res. 2019, 28, 1529–1546. [Google Scholar] [CrossRef]
  102. Zavala, D.Q.; Díaz, J.R.M.; Melo, J.L.A.; Pliego, R.G.; Vázquez, H.A.J.; Ferrara, J.G.T.; Guevara-Salazar, J.A. Physicochemical interpretation, with QSAR/SAR analysis, of how the barriers of Pseudomonas aeruginosa bacteria were penetrated by para-substituted N-arylbenzylimines: Synthesis, characterization, and in vitro antibacterial effect. J. Mex. Chem. Soc. 2021, 65, 376–395. [Google Scholar] [CrossRef]
  103. Hansch, C.; Leo, A. Substituent Constants for Correlation Analysis in Chemistry and Biology, 7th ed.; John Wiley & Sons: Hoboken, NJ, USA, 1979; 339p. [Google Scholar] [CrossRef]
  104. Williford, C.; Stevens, E. Strain energies as a steric descriptor in QSAR calculations. QSAR Comb. Sci. 2004, 23, 495–505. [Google Scholar] [CrossRef]
  105. Alipour, M.; Safari, Z. From information theory to quantitative description of steric effects. Phys. Chem. Chem. Phys. 2016, 18, 17917–17929. [Google Scholar] [CrossRef]
  106. Teixeira, J.; Gaspar, A.; Garrido, E.M.; Garrido, J.; Borges, F. Hydroxycinnamic Acid Antioxidants: An Electrochemical Overview. BioMed Res. Int. 2013, 2013, 251754. [Google Scholar] [CrossRef]
Figure 1. BHX, tubacin, trichostatin A (TSA), and suberoylanilide hydroxamic acid (SAHA).
Figure 1. BHX, tubacin, trichostatin A (TSA), and suberoylanilide hydroxamic acid (SAHA).
Pharmaceuticals 15 00690 g001
Figure 2. (a) Steric effects and noncovalent interactions considered as structural design elements of compounds 2ai. (b) The expected method for compounds 2ai to cap the entrance of the active site cleft through noncovalent binding interactions in the shallow L1 and L2 pockets.
Figure 2. (a) Steric effects and noncovalent interactions considered as structural design elements of compounds 2ai. (b) The expected method for compounds 2ai to cap the entrance of the active site cleft through noncovalent binding interactions in the shallow L1 and L2 pockets.
Pharmaceuticals 15 00690 g002
Scheme 1. Synthesis of DPCH derivatives 2ai.
Scheme 1. Synthesis of DPCH derivatives 2ai.
Pharmaceuticals 15 00690 sch001
Scheme 2. Mechanistic proposal for the conversion of 1 to 2. The structure of the enol intermediate B is shown in the right panel.
Scheme 2. Mechanistic proposal for the conversion of 1 to 2. The structure of the enol intermediate B is shown in the right panel.
Pharmaceuticals 15 00690 sch002
Figure 3. (a) Ortep plot at 50% probability level of the DMSO (not shown) solvate of compound 2a-(4R, 5R) enantiomer. (b) Stick plot of compound 2a—view through the 4,5-dihydro-pyrazole ring—to appreciate the steric effect of the four phenyl rings and the cis disposition between H4 and H5.
Figure 3. (a) Ortep plot at 50% probability level of the DMSO (not shown) solvate of compound 2a-(4R, 5R) enantiomer. (b) Stick plot of compound 2a—view through the 4,5-dihydro-pyrazole ring—to appreciate the steric effect of the four phenyl rings and the cis disposition between H4 and H5.
Pharmaceuticals 15 00690 g003
Figure 4. Binding poses at DD2-HDAC6 catalytic domain obtained through blind docking of the target compounds. Compounds: (a) 2a, (b) 2b, and (c) 2e; Zn2+ is depicted as yellow sphere. In the left panel, HDAC6 is depicted in white cartoon; in the middle panel, a zoom of the catalytic domain is shown where the AAR are as sticks, and the ligands are shown using a ball and stick representation; in the right panel, a surface representation of the catalytic tunnel is depicted in gray where the insertion of the N-Ph (a), C5-PhOH (b), and C3-Ph (c) rings are appreciated. Figures were built with Pymol and UCSF Chimera software.
Figure 4. Binding poses at DD2-HDAC6 catalytic domain obtained through blind docking of the target compounds. Compounds: (a) 2a, (b) 2b, and (c) 2e; Zn2+ is depicted as yellow sphere. In the left panel, HDAC6 is depicted in white cartoon; in the middle panel, a zoom of the catalytic domain is shown where the AAR are as sticks, and the ligands are shown using a ball and stick representation; in the right panel, a surface representation of the catalytic tunnel is depicted in gray where the insertion of the N-Ph (a), C5-PhOH (b), and C3-Ph (c) rings are appreciated. Figures were built with Pymol and UCSF Chimera software.
Pharmaceuticals 15 00690 g004
Figure 5. In vitro wound closure cell migration assay with the BC cell line MDA-MB-231 at 0, 24, and 48 h of incubation. (a) Control, (b) compound 2b, and (c) compound 2c at 15 µM concentration. (d) Dose-dependent curve of HDAC6 inhibition by compound 2b. Data represent mean ± SEM [* significantly different from AA (p < 0.05)]. (e) Confocal microscopy images, MCF-7 cells exposed to compound 2b at a concentration of 10 µM: increased resolution of lasser (left) and low-resolution image (right).
Figure 5. In vitro wound closure cell migration assay with the BC cell line MDA-MB-231 at 0, 24, and 48 h of incubation. (a) Control, (b) compound 2b, and (c) compound 2c at 15 µM concentration. (d) Dose-dependent curve of HDAC6 inhibition by compound 2b. Data represent mean ± SEM [* significantly different from AA (p < 0.05)]. (e) Confocal microscopy images, MCF-7 cells exposed to compound 2b at a concentration of 10 µM: increased resolution of lasser (left) and low-resolution image (right).
Pharmaceuticals 15 00690 g005
Figure 6. Quantitative structure–activity relationship (QSAR) between the pIC50 (M−1) and Es values of DPCH derivatives on BC cellular lines (a) MCF-7 and (b) MDA-MB-231. QSAR between the pIC50 (M−1), (c) π, and (d) σH values of DPCH derivatives on healthy cellular line MCF-10A. Polynomial regression of second order through one-way ANOVA test and constant values were analyzed by Student’s t-test: (a) pIC50 = -638.5 ± 1.6 ES2 + 18.6 ± 0.7 ES + 4.46 ± 0.04 (n = 9, p < 0.05, r = 0.9266); (b) pIC50 = -831 ± 36 ES2 + 30.4 ± 1.3 ES + 4.30 ± 0.02 (n = 9, p < 0.001, r = 0.9634); (c) pIC50 = 1.50 ± 0.15 π2–0.076 ± 0.011 + 3.969 ± 0.002 (n = 9, p < 0.01, r = 0.9584); (d) 6.4 ± 0.6 σH2 + 1.20 ± 0.12 σH + 3.993 ± 0.002 (n = 6, p < 0.05, r = 0.9186). The p < 0.05 and p < 0.01 values represent statistically significant differences at 95.0% and 99.0% confidence, respectively.
Figure 6. Quantitative structure–activity relationship (QSAR) between the pIC50 (M−1) and Es values of DPCH derivatives on BC cellular lines (a) MCF-7 and (b) MDA-MB-231. QSAR between the pIC50 (M−1), (c) π, and (d) σH values of DPCH derivatives on healthy cellular line MCF-10A. Polynomial regression of second order through one-way ANOVA test and constant values were analyzed by Student’s t-test: (a) pIC50 = -638.5 ± 1.6 ES2 + 18.6 ± 0.7 ES + 4.46 ± 0.04 (n = 9, p < 0.05, r = 0.9266); (b) pIC50 = -831 ± 36 ES2 + 30.4 ± 1.3 ES + 4.30 ± 0.02 (n = 9, p < 0.001, r = 0.9634); (c) pIC50 = 1.50 ± 0.15 π2–0.076 ± 0.011 + 3.969 ± 0.002 (n = 9, p < 0.01, r = 0.9584); (d) 6.4 ± 0.6 σH2 + 1.20 ± 0.12 σH + 3.993 ± 0.002 (n = 6, p < 0.05, r = 0.9186). The p < 0.05 and p < 0.01 values represent statistically significant differences at 95.0% and 99.0% confidence, respectively.
Pharmaceuticals 15 00690 g006
Figure 7. (a) DPPH radical scavenging activity of compounds 2ai vs ascorbic acid (AA) at 100 µM. Data represent mean ± SEM [* significantly different from AA (p < 0.05)]. The assays were performed in triplicate. Antioxidant IC50 values (μM): 38 ± 3 (2a), 26 ± 2 (2b), 16 ± 4 (2c), 32 ± 3 (2d), 26 ± 2 (2e), 35 ± 3 (2f), 23 ± 3 (2g), 17 ± 3 (2h), 21 ± 2 (2i), and 13 ± 2 (AA). (b) Comparative graph between the antioxidant and antiproliferative (MCF-10A, MCF-7 and MDA-MB231 cell lines) IC50 values (radial axis) of compounds 2ai (periphery).
Figure 7. (a) DPPH radical scavenging activity of compounds 2ai vs ascorbic acid (AA) at 100 µM. Data represent mean ± SEM [* significantly different from AA (p < 0.05)]. The assays were performed in triplicate. Antioxidant IC50 values (μM): 38 ± 3 (2a), 26 ± 2 (2b), 16 ± 4 (2c), 32 ± 3 (2d), 26 ± 2 (2e), 35 ± 3 (2f), 23 ± 3 (2g), 17 ± 3 (2h), 21 ± 2 (2i), and 13 ± 2 (AA). (b) Comparative graph between the antioxidant and antiproliferative (MCF-10A, MCF-7 and MDA-MB231 cell lines) IC50 values (radial axis) of compounds 2ai (periphery).
Pharmaceuticals 15 00690 g007
Table 1. Physicochemical properties of 2ai and prediction of oral absorption (%ABS) compared with the reference compounds (tubacin, TSA, and SAHA).
Table 1. Physicochemical properties of 2ai and prediction of oral absorption (%ABS) compared with the reference compounds (tubacin, TSA, and SAHA).
Comp.MWLogPHAHDRBVLRLogSTPSA
2]
MR
[cm3/mol−1]
%ABS
2a448.522.76360−5.077.0132.182
2b482.973.36360−5.7377.0136.782
2c478.552.77370−5.0186.2139.480
2d527.423.46361−5.8377.0139.882
2e464.522.27460−4.7097.2133.976
2f464.522.27460−4.7097.2133.976
2g464.522.27460−4.7097.2133.9176
2h557.443.07371−5.8586.2147.080
2i492.572.67380−5.3186.2144.280
Tubacin721.877.2104163−9.41168.4200.051
TSA302.372.25260−3.2669.691.785
SAHA264.322.35380−3.3378.470.682
Abbreviations: MW = molecular weight (g mol−1); LogP = logarithm of octanol–water partition coefficient; HA = hydrogen acceptors; HD = hydrogen donors; RB = rotatable bonds; VLR = violations of Lipinski’s rules; LogS = logarithm of the solubility in water; TPSA = topological polar surface area; %ABS = absorption percentage; NA = not applicable.
Table 2. Cytotoxic activity of 2ai and reference compounds (IC50 µM) on several cell lines.
Table 2. Cytotoxic activity of 2ai and reference compounds (IC50 µM) on several cell lines.
CompoundsMCF-7MDA-MB-2313T3/NIHMCF-10A
2a41 ± 560 ± 2>100>100
2b35 ± 324 ± 219 ± 233 ± 4
2c26 ± 133 ± 3>100>100
2d24 ± 226 ± 125 ± 322 ± 3
2e24 ± 332 ± 332 ± 255 ± 5
2f28 ± 233 ± 236 ± 245 ± 4
2g23 ± 133 ± 328 ± 230 ± 3
2h101 ± 2108 ± 1103 ± 2>100
2i71 ± 564 ± 4>100>100
TSA0.5 ± 20.4 ± 11 ± 20.2 ± 0.06
SAHA8 ± 12.5 ± 110 ± 212 ± 1.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Balbuena-Rebolledo, I.; Rivera-Antonio, A.M.; Sixto-López, Y.; Correa-Basurto, J.; Rosales-Hernández, M.C.; Mendieta-Wejebe, J.E.; Martínez-Martínez, F.J.; Olivares-Corichi, I.M.; García-Sánchez, J.R.; Guevara-Salazar, J.A.; et al. Dihydropyrazole-Carbohydrazide Derivatives with Dual Activity as Antioxidant and Anti-Proliferative Drugs on Breast Cancer Targeting the HDAC6. Pharmaceuticals 2022, 15, 690. https://doi.org/10.3390/ph15060690

AMA Style

Balbuena-Rebolledo I, Rivera-Antonio AM, Sixto-López Y, Correa-Basurto J, Rosales-Hernández MC, Mendieta-Wejebe JE, Martínez-Martínez FJ, Olivares-Corichi IM, García-Sánchez JR, Guevara-Salazar JA, et al. Dihydropyrazole-Carbohydrazide Derivatives with Dual Activity as Antioxidant and Anti-Proliferative Drugs on Breast Cancer Targeting the HDAC6. Pharmaceuticals. 2022; 15(6):690. https://doi.org/10.3390/ph15060690

Chicago/Turabian Style

Balbuena-Rebolledo, Irving, Astrid M. Rivera-Antonio, Yudibeth Sixto-López, José Correa-Basurto, Martha C. Rosales-Hernández, Jessica Elena Mendieta-Wejebe, Francisco J. Martínez-Martínez, Ivonne María Olivares-Corichi, José Rubén García-Sánchez, Juan Alberto Guevara-Salazar, and et al. 2022. "Dihydropyrazole-Carbohydrazide Derivatives with Dual Activity as Antioxidant and Anti-Proliferative Drugs on Breast Cancer Targeting the HDAC6" Pharmaceuticals 15, no. 6: 690. https://doi.org/10.3390/ph15060690

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