Next Article in Journal
Dietary Betaine Attenuates High-Carbohydrate-Diet-Induced Oxidative Stress, Endoplasmic Reticulum Stress, and Apoptosis in Mandarin Fish (Siniperca chuatsi)
Previous Article in Journal
The Inhibition of Neuropathic Pain Incited by Nerve Injury and Accompanying Mood Disorders by New Heme Oxygenase-1 Inducers: Mechanisms Implicated
Previous Article in Special Issue
Exploring the Antioxidant Properties of Caffeoylquinic and Feruloylquinic Acids: A Computational Study on Hydroperoxyl Radical Scavenging and Xanthine Oxidase Inhibition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Coumarin N-Acylhydrazone Derivatives: Green Synthesis and Antioxidant Potential—Experimental and Theoretical Study

by
Dušica M. Simijonović
1,
Dejan A. Milenković
1,*,
Edina H. Avdović
1,
Žiko B. Milanović
1,
Marko R. Antonijević
1,
Ana D. Amić
2,
Zana Dolićanin
3 and
Zoran S. Marković
1,4,*
1
Department of Science, Institute for Information Technologies, University of Kragujevac, Jovana Cvijića bb, 34000 Kragujevac, Serbia
2
Department of Chemistry, Josip Juraj Strossmayer University of Osijek, Ulica Cara Hadrijana 8A, 31000 Osijek, Croatia
3
Department of Biomedical Sciences, State University of Novi Pazar, Vuka Karadžića bb, 36300 Novi Pazar, Serbia
4
Department of Natural Science and Mathematics, State University of Novi Pazar, Vuka Karadžića bb, 36300 Novi Pazar, Serbia
*
Authors to whom correspondence should be addressed.
Antioxidants 2023, 12(10), 1858; https://doi.org/10.3390/antiox12101858
Submission received: 19 September 2023 / Revised: 4 October 2023 / Accepted: 8 October 2023 / Published: 13 October 2023
(This article belongs to the Special Issue Theoretical and Computational Chemistry in Antioxidant Research)

Abstract

:
Coumarin N-acylhydrazone derivatives were synthesized in the reaction of 3-acetylcoumarin and different benzohydrazides in the presence of molecular iodine as catalyst and at room temperature. All reactions were rapidly completed, and products were obtained in good to excellent yields. It is important to emphasize that four products were reported for the first time in this study. The obtained compounds were subjected to evaluation of their in vitro antioxidative activity using DPPH, ABTS, and FRAP methods. It was shown that products with a catechol moiety in their structure are the most potent antioxidant agents. The thermodynamic parameters and Gibbs free energies of reactions were used to determine the most probable mechanism of action. The results of in silico examination emphasize the need to take solvent polarity and free radical species into account when examining antiradical action. It was discovered by using computational approaches that HAT and SPLET are competitive molecular pathways for the radical scavenging activity of all compounds in polar mediums, while the HAT is the dominant mechanism in non-polar environments.

1. Introduction

Coumarins constitute a large group of natural compounds that are extensively prevalent in nature [1,2]. In addition to fruits, they are most commonly found in the roots, stems, and leaves of higher plants. Some essential oils, such as lavender oil, cinnamon bark oil, and cassia leaf oil, contain large amounts of different coumarin derivatives. Certain varieties of cinnamon are the most common source of coumarins for humans [3]. Coumarins are relatively simple compounds that basically consist of fused α-pyrone and a benzene ring. Coumarins that occur naturally exhibit a wide range of pharmacological activities and are frequently employed by researchers in the design and development of novel synthetic coumarins [4]. The significance of coumarin hybrids in the fields of medicine and pharmacy is noteworthy due to their relatively low level of adverse effects on living things, as reported in the literature [5]. Several pharmacological activities, such as antioxidant [6,7,8], anticancer [9,10], antitubercular [11], antibacterial [12,13,14], antifungal [15], antiviral [16], and anticoagulant [17,18,19,20,21], have been exhibited by many of the obtained hybrid compounds. Chemically modified coumarins have great potential as drugs for the treatment of various diseases. Several coumarin derivatives, including warfarin (I), dicoumarol (II), phenprocoumon (III), armilarizine A (IV), carbochromen (V) and chlorochromen (VI), have found significant clinical application, Figure 1.
Hydrazone compounds are a well-known group of compounds with specific structural units (R2C=N–NH2). These compounds are of great significance in organic synthesis and in the creation of compounds that exhibit a wide range of biological activities. The literature describes the coumarin compounds with various hydrazide–hydrazone pharmacophores attached to position three in the chromen ring [22,23,24,25,26]. The synthesis of coumarin hydrazone derivatives involves the condensation of diverse 3-formylcoumarin with different hydrazides in the presence of acid as a catalyst [22,23]. According to the literature, a method is known for the synthesis of a series of 3-acylhydrazono-4-hydroxycoumarins using the reaction of 3-acetyl-4-hydroxycoumarin with the corresponding hydrazides [24,25,26]. However, it is worth pointing out that the reactions between 3-acetylcoumarin and various benzoyl hydrazides have been described in only one article [27]. Pangal et al. conducted a study wherein they successfully synthesized three derivatives of coumarin N-acylhydrazone through the reaction of 3-acetylcoumarin with appropriate hydrazides under reflux in methanol.
In this paper, we report a green methodology for the synthesis of coumarin N-acylhydrazone derivatives in ethanol and in the presence of molecular iodine as an efficient, relatively nontoxic, and inexpensive catalyst. Additionally, this study involves the assessment of the antioxidative potential of the obtained products by using both experimental and theoretical approaches.

2. Materials and Methods

All chemicals used for the synthesis were purchased from Sigma Aldrich, Darmstadt, Germany with purities above 98%. The IR spectra were obtained on a PerkinElmer Spectrum One FT-IR spectrometer using the KBr disc. The NMR spectra were recorded on a Varian Gemini spectrometer (200 MHz for 1H and 50 MHz for 13C) by using DMSO-d6 as solvent. All UV-Vis determinations were carried out using a PerkinElmer, Lambda 365 UV/Vis Spectrophotometer. Elemental microanalysis for carbon, hydrogen, and nitrogen was performed at the Institute for Information Technologies, University of Kragujevac. Elemental (C, H, N) analysis of the samples was carried out on an Elemental analysis system NC Technologies, ECS 8020 CHNOS, model-dual furnace-NC.

2.1. Synthesis of Coumarin N-Acylhydrazone Derivatives

Coumarin N-acylhydrazone derivatives 3 were obtained in the reaction of 3-acetylcoumarin (0.001 mol) and corresponding benzohydrazide 2 (0.001 mol) in presence of molecular iodine as catalyst (10 mol%) in absolute ethanol at room temperature. The progress of the reaction was monitored using thin-layer chromatography (TLC). Upon completion of the reaction, the purified product was isolated by filtration. All coumarin N-acylhydrazone derivatives (3af) were characterized with 1H NMR, 13C NMR, IR, and UV-Vis spectra. In addition, elemental analysis confirmed the purity of the newly synthesized compounds 3bf.
(E)-4-hydroxy-N′-(1-(2-oxo-2H-chromen-3-yl)ethylidene)benzohydrazide (3a): 1H NMR (200 MHz, DMSO-d6) δ 10.57 (s, 1H), 10.11 (s, 1H), 8.22 (s, 1H), 7.88 (dd, J = 7.8, 1.5 Hz, 1H), 7.83–7.74 (m, 2H), 7.66 (ddd, J = 8.7, 7.2, 1.7 Hz, 1H), 7.50–7.34 (m, 2H), 6.91–6.80 (m, 2H), 2.32 (s, 3H); 13C NMR (50 MHz, DMSO-d6) δ 160.7, 159.3, 153.5, 141.5, 132.4, 130.3, 129.2, 127.0, 124.8, 124.7, 118.9, 116.1, 114.9, 16.7; IR (KBr): 3282 (OH), 3205 (NH), 3040 (ArC-H), 3018 (CH), 1707, 1659 (C=O), 1276 (C–N), 1222, 1163 (C–O) cm−1; UV (λmax): 269, 327 nm.
(E)-2-hydroxy-N′-(1-(2-oxo-2H-chromen-3-yl)ethylidene)benzohydrazide (3b): 1H NMR (200 MHz, DMSO-d6) δ 11.79 (s, 1H), 11.41 (s, 1H), 8.28 (s, 1H), 7.94 (dd, J = 20.2, 7.8 Hz, 2H), 7.67 (m, J = 8.7, 7.3, 1.6 Hz, 1H), 7.46 (d, J = 8.4 Hz, 2H), 7.38 (d, J = 7.7 Hz, 1H), 7.07–6.85 (m, 2H), 2.30 (s, 3H); 13C NMR (50 MHz, DMSO-d6) δ 159.3, 156.5, 153.5, 141.7, 133.5, 132.5, 130.7, 129.3, 124.9, 119.7, 118.8, 116.9, 116.1, 15.8; IR (KBr): 3435 (OH), 3262 (NH), 3051 (ArC-H), 2928 (CH), 1725, 1691 (C=O), 1299 (C–N), 1161 (C–O) cm−1; UV (λmax): 267, 330 nm.
(E)-3,4-dihydroxy-N′-(1-(2-oxo-2H-chromen-3-yl)ethylidene)benzohydrazide (3c): 1H NMR (200 MHz, DMSO-d6) δ 10.53 (s, 1H), 9.61 (s, 1H), 9.29 (s, 1H), 8.22 (s, 1H), 7.95–7.55 (m, 2H), 7.52–7.12 (m, 4H), 6.82 (d, J = 8.1 Hz, 1H), 2.31 (s, 3H); 13C NMR (50 MHz, DMSO-d6) δ 159.3, 153.5, 149.1, 145.0, 141.5, 132.4, 129.2, 124.8, 124.7, 120.3, 118.9, 116.1, 115.0, 16.2; IR (KBr): 3429 (OH, NH), 3077 (ArC-H), 2929 (CH), 1712, 1698 (C=O), 1309 (C–N), 1220 (C–O) cm−1; UV (λmax): 271, 328 nm; C18H14N2O5 (FW = 338.32): C, 63.90; N, 8.28; H, 4.17%; found: C, 64.27; N, 8.15; H, 4.52%.
(E)-2,3-dihydroxy-N′-(1-(2-oxo-2H-chromen-3-yl)ethylidene)benzohydrazide (3d): 1H NMR (200 MHz, DMSO-d6) δ 11.46 (s, 1H), 10.88 (s, 1H), 9.86 (s, 1H), 8.27 (s, 1H), 7.98–7.58 (m, 2H), 7.52–7.15 (m, 3H), 7.08–6.59 (m, 2H), 2.28 (s, 3H); 13C NMR (50 MHz, DMSO-d6) δ 159.3, 153.5, 146.1, 141.7, 132.6, 129.3, 124.9, 120.2, 119.3, 118.9, 118.8, 116.1, 15.9; IR (KBr): 3502, 3430 (OH), 3250 (NH), 3064 (ArC-H), 2873 (CH), 1690, 1649 (C=O), 1283 (C–N), 1140 (C–O) cm−1; UV (λmax): 266, 333 nm; C18H14N2O5 (FW = 338.32): C, 63.90; N, 8.28; H, 4.17%; found: C, 63.67; N, 8.35; H, 4.56%.
(E)-3,4,5-trimethoxy-N′-(1-(2-oxo-2H-chromen-3-yl)ethylidene)benzohydrazide (3e): 1H NMR (200 MHz, DMSO-d6) δ 10.75 (s, 1H), 8.24 (s, 1H), 7.88 (d, J = 7.3 Hz, 1H), 7.73–7.59 (m, 1H), 7.41 (dd, J = 15.3, 7.9 Hz, 2H), 7.20 (s, 2H), 3.86 (s, 6H), 3.73 (s, 3H), 2.34 (s, 3H); 13C NMR (50 MHz, DMSO-d6) δ 159.2, 153.5, 152.6, 132.5, 129.3, 129.0, 126.9, 124.8, 118.82, 116.1, 105.9, 60.2, 16.5; IR (KBr): 3428 (OH), 3277 (NH), 3048 (ArC-H), 2910 (CH), 1719, 1665 (C=O), 1230 (C–N), 1127 (C–O) cm−1; UV (λmax): 276, 329 nm; C21H20N2O6 (FW = 396.40): C, 63.63; N, 7.07; H, 5.09%; found: C, 63.87; N, 6.85; H, 4.72%.
(E)-4-hydroxy-3-methoxy-N′-(1-(2-oxo-2H-chromen-3-yl)ethylidene)benzohydrazide (3f): 1H NMR (200 MHz, DMSO-d6) δ 10.59 (s, 1H), 9.73 (s, 1H), 8.22 (s, 1H), 7.87 (dd, J = 7.7, 1.4 Hz, 1H), 7.66 (ddd, J = 8.8, 7.3, 1.6 Hz, 1H), 7.51–7.33 (m, 4H), 6.87 (d, J = 8.3 Hz, 1H), 3.85 (s, 3H), 2.33 (s, 3H); 13C NMR (50 MHz, DMSO-d6) δ 159.3, 153.5, 150.2, 141.5, 132.4, 129.2, 127.0, 124.8, 124.5, 122.1, 118.9, 116.1, 114.9, 55.9, 16.4; IR (KBr): 3428 (OH), 3337 (NH), 3078 (ArC-H), 2941 (CH), 1712, 1649 (C=O), 1280 (C–N), 1121 (C–O) cm−1; UV (λmax): 269, 324 nm; C19H16N2O5 (FW = 352.35): C, 64.77; N, 7.95; H, 4.58%; found: C, 64.97; N, 7.79; H, 4.53%.

2.2. DPPH Radical Scavenging Assay

The free radical scavenging activities of referent compounds and synthesized products were determined using a 1,1-diphenyl-2-picryl-hydrazyl (DPPH) assay [28]. The tested samples (20 µL of different concentrations dissolved in DMSO and 980 µL of methanol) were mixed with the same volume of the solution of DPPH in methanol (0.05 mM). The prepared samples were shaken well and left at room temperature in the dark for 20 min and 60 min. After the incubation period, absorbance was determined at 517 nm by using the methanol as a blank control. All tests were run in triplicate and averaged. The results are presented as mean value ± standard deviation. Nordihydroguaiaretic acid (NDGA) and quercetin were used as positive controls. For products that exert good activity, IC50 values were determined, and the corresponding concentrations used for this determination are given in Table S1. The stoichiometric factor (SF) was calculated using next equation:
S F = D P P H 0 2 × I C 50 .

2.3. ABTS Radical Cation Scavenging Assay

The antioxidant capacity was measured using the 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) diammonium salt) (ABTS) assay according to the modified procedure of Pontiki et al. [29]. In this assay, stock solutions of ABTS (7 mM) and potassium persulfate (2.45 mM) were first prepared. The working solution was then prepared by mixing the two stock solutions in equal quantities and allowing them to react for 12–16 h at room temperature in the dark. The ABTS+• radical cation is diluted with methanol to obtain the solution with an absorbance of 0.70 units at 734 nm. Different concentrations of samples were prepared in DMSO. Next, 20 μL of sample and 980 μL of methanol were mixed, then an equal amount of ABTS was added and the absorbance was measured at 734 nm. IC50 values were determined for products with noticeable activity and the concentrations used for this determination are given in Table S2.

2.4. Ferric Ion Reducing Capacity Assay

The ferric ion reducing capacity (FRAP) assay is a direct method for measuring the combined antioxidant activity of reductive antioxidants in a substance under investigation. In this study, the FRAP assay was conducted with some modifications to the method described by Pownall et al. [30]. The stock solutions of tested compounds were prepared in DMSO, and then sample dilutions were prepared in phosphate buffer, pH 7.4. After that, 500 µL of diluted samples were mixed with 250 µL of 1% potassium ferricyanide solution followed by incubation for 20 min at 50 °C. Following the incubation period, a mixture was prepared by combining 500 µL of the sample with 500 µL of 10% trichloroacetic acid, 100 µL of 0.1% ferric chloride, and 500 µL of deionized water. The mixture was subjected to an incubation period of 10 min at room temperature, after which, the absorbance was measured at the wavelength of 700 nm. Concentrations of investigated compounds and referent standard ascorbic acid in the sample were 10 µM. The results of reducing the power of compounds were expressed as % Activity of Ascorbic acid (% AAa) [31].
%   A A a = A s a m p l e A r e f e r e n c e × 100

2.5. Computational Methodology

The Gaussian09 software package was used to carry out all calculations [32]. Quantum chemical calculations based on density functional theory, specifically M06-2X functional in conjunction with 6-311++G(d,p) basis set (with polarization and diffuse functions included), were used to optimize the structures of parent molecules as well as of the corresponding radicals, anions, and radical cations [33,34]. Previous research has demonstrated that the M062-X method is suitable for thermodynamic and kinetic investigation of reaction mechanisms of examined compounds with free radicals and that it describes short- and medium-range interatomic interactions well (500 pm) [35]. The SMD continuum solvation model [36] was used to simulate the structures in benzene (ε = 2.27) and methanol (ε = 32.61) with no geometrical restrictions. These solvents were chosen to replicate both polar and nonpolar settings as well as the environment in which experimental measurements were conducted.
For the assessment of the antioxidant activity, three main radical scavenging mechanisms—Hydrogen Atom Transfer (HAT), Single-Electron Transfer Followed by Proton Transfer (SET-PT), and Sequential Proton Loss-Electron Transfer (SPLET) were used [37,38,39,40]. In the first reaction mechanism, the O–H bond is homolytic cleavage, resulting in the separation of a hydrogen atom and a radical species (A–O) (Equation (1)). In the first step of the SET-PT mechanism, a molecule loses an electron, which leads to the formation of the radical cation (A–OH+•), which loses a proton in the second step (Equations (2) and (3)). Another two-step mechanism is the SPLET mechanism. The antioxidant molecule is used to create an anion (A–O) in the first stage, and the matching radical is created in the second step following an electron transfer (Equations (4) and (5)).
A–OH → A–O + H
A–OH → A–OH+• + e
A–OH+• → A–O + H+
A–OH → A–O + H+
A–O → A–O + e
BDE (Bond Dissociation Enthalpy, Equation (6)), IP (Ionization Potential, Equation (7)), and PDE (Proton Dissociation Enthalpy, Equation (8)), as well as PA (Proton Affinity, Equation (9)) and ETE (Electron Transfer Enthalpy, Equation (10)), are the thermodynamic parameters driving the antioxidative mechanisms. The following equations were used to calculate the parameters, at 298.15 K, from the enthalpies of the optimized species [41,42].
BDE = H(A–O) + H(H) − H(A–OH)
IP = H(A–OH•+) + H(e) − H(A–OH)
PDE = H(A–O) + H(H+) − H(A–OH•+)
PA = H(A–O) + H(H+) − H(A–OH)
ETE = H(A–O) + H(e) − H(A–O)
The solvated electron and proton enthalpies for the M062-X technique in benzene (−877.4 and −10.5 kJ mol−1) and methanol (−1065.4 and −61.4 kJ mol−1) were obtained from the available data in the literature [43]. The Natural Bond Orbital (NBO) method and NBO 6.0 package were used to conduct the population study [44]. The intermolecular orbitals’ role in the examined compounds’ anions and radicals is highlighted using NBO analysis.

3. Results and Discussion

Our recent results related to the preparation of coumarin-hydroxybenzohydrazide derivatives prompted us to carry out the synthesis of structurally similar compounds starting from 3-acetylcoumarin [26]. The synthesis of various coumarin hydrazones (3) from 3-acetylcoumarin (1) and the corresponding benzoyl hydrazide (2) was presented in Scheme 1. The reaction between 1 and 4-hydroxybenzohydrazide (2a) in ethanol was used as a model for optimizing reaction conditions (Table 1). Initially, the reaction was conducted at room temperature and without a catalyst, but only trace amounts of the reaction product were obtained. The moderate yield of the product was observed after 10 h of heating the reaction mixture under reflux. In order to attain a higher product yield, acetic acid was used as a catalyst. In this instance, an adequate yield of 3a was obtained by increasing the catalyst concentration to 20 mol% and heating the mixture for 10 h. However, the catalyst had a substantial effect on both the reaction rate and yield. Specifically, the use of molecular iodine as a catalyst allowed the desired product 3a to be obtained in only 20 min and without heating (Table 1). The optimal conditions for carrying out these reactions were 10 mol% I2 as a catalyst, room temperature, and 20 min.
The results indicate that the applied reaction conditions result in the outstanding dissolution of the reactants, rapid precipitation of the product, and simple isolation of the reaction products. Table 2 shows that all products were obtained with yields ranging from outstanding to very good. All reactions were completed very rapidly, except for compounds 3d and 3e, for which a slight extension of the reaction time was required. It is essential to note that four obtained products have been reported for the first time in this study (3cf). The isolated products’ structures were characterized using NMR, IR, and UV-Vis spectroscopies. Additionally, elemental analysis was performed on newly synthesized compounds.

3.1. Structural Characterization of Coumarin N-Acylhydrazone Derivatives

The signals in the 1H NMR spectra of derivatives 3 were assigned to three groups of protons. Singlets originating from the protons of the methyl group were observed at approximately 2.30 ppm in the 1H NMR spectra of all the compounds. The aromatic protons of coumarin and hydrazide moiety were in a region of 6.64–8.01 ppm. The hydrogen atoms of NH and phenolic OH groups were observed in the regions of 10.53–11.79 ppm and 9.29–11.41 ppm as broad singlets. In the spectra of compounds 3e and 3f, singlets originating from protons of methoxy groups attached to the aromatic ring at about 3.85 ppm were observed. In the 13C NMR spectra of all the products, the carbon atom of the methyl group appears around 17 ppm. Signals in the range of 105–145 ppm were attributed to aromatic carbon atoms. Peaks detected at around 150, 153, and 160 ppm originate from C–O, C–N, and C=O carbon atoms. The additional peaks at about 60 ppm were observed in the spectra of compounds 3e and 3f because of the presence of methoxy groups.
The vibration of OH and NH groups corresponds to regions at approximately 3400 and 3250 cm−1 in the IR spectra of investigated compounds. The peaks in region 1650–1725 cm−1 are assigned to the C=O of hydrazone group and lactone ring, while bands present at lower values are attributed to the stretching vibration of C–N, C–O, and aromatic rings. In the UV-Vis spectra of all the compounds, two major absorption bands appear around 270 and 330 nm.

3.2. The Proposed Mechanism for the Synthesis of Coumarin N-Acylhydrazone Derivatives 3

The suggested mechanism for the synthesis of coumarin N-acylhydrazone derivatives 3 was presented in Scheme 2. In this reaction, molecular iodine most likely acts only as a precatalytic. In the first step, as a solvent, ethanol reacts with molecular iodine to give the Brønsted acid HI, which further acts as a catalyst, as shown in Scheme 2. In the subsequent step, the carbonyl group of 3-acetylcoumarin is protonated with Brønsted acid HI, followed by the nucleophilic attack of hydrazide’s nitrogen atom to the carbonyl group, and the deprotonation of the same nitrogen atom. In this manner, an appropriate hemiaminal is formed as an intermediate. The resulting intermediate undergoes dehydration in a similar way. During this stage, the HI species facilitates the protonation of the hydroxyl group, leading to the formation of a water molecule, with simultaneous deprotonation of the nitrogen atom. In this way, corresponding coumarin N-acylhydrazone derivative 3 is formed.

3.3. Antioxidant Activity of Coumarin N-Acylhydrazone Derivatives

The DPPH and ABTS scavenging ability, as well as the ferric ion reducing capacity, of the synthesized coumarin N-acylhydrazone derivatives (3af) were evaluated to determine their in vitro antioxidant properties. The results are presented in Table 3 and Table 4. In this study, NDGA, quercetin, (S)-6-methoxy-2,5,7,8-tetramethylchromane-2-carboxylic acid (Trolox), and ascorbic acid were used as referent compounds.
The results obtained from the DPPH assay revealed that compounds 3c and 3d expressed the best antioxidant activity with IC50 values of 2.1 µM and 3.2 µM. The activity of these compounds is slightly lower than the activity of the positive controls, quercetin and NDGA. In order to determine whether the most active compounds belong to the group of good antioxidants, the SF was determined [26,45]. It is known that compounds with an SF greater than 2 are good antioxidants. Namely, the tested compounds 3c and 3d, with SF values 6 and 3.9, belong to the group of excellent antioxidants. Other tested compounds showed extremely low activity toward the DPPH radical even at significantly higher concentrations, Table 3.
The ABTS assay, which is a commonly employed technique for evaluating in vitro antioxidant capacity, was utilized to examine the radical scavenging potential of the investigated compounds, as presented in Table 4. The results obtained from this study indicate that compounds 3c and 3d exhibit significant scavenging activity against ABTS radical cation, with IC50 values of 2.2 µM and 3.0 µM, respectively. This finding is consistent with previous DPPH assay results. Compound 3f exhibited better activity relative to the DPPH results, exhibiting an efficacy of roughly 50% at a concentration of 100 µM.
The results of the FRAP assay revealed that all the examined compounds exhibited higher capacities for Fe3+ reduction than ascorbic acid, which was used as the standard. It should be emphasized that the obtained results are in good agreement with the activities of these compounds against DPPH and ABTS radicals. Namely, compound 3c showed the greatest reducing ability, followed by compounds 3d and 3f. All the obtained results point out the compounds 3c and 3d, with two neighboring hydroxyl groups, i.e., compounds with a catechol fragment, as the best antioxidants. In addition, moderate activity against the ABTS radical cation is observed for compound 3f with vanillic fragment. The reason for this is the resonance and electron donating effects of these groups, as well as the very effective stabilization of the formed phenoxy radical by an intramolecular hydrogen bond.

3.4. Determination of the Plausible Mechanisms

The optimized geometries of the newly synthesized compounds, as shown in Figure 2, were obtained using the M06-2X/6-311++G(d,p) theoretical model. The geometries of the compounds under investigation exhibited a correlation with the geometries of structurally analogous compounds obtained in previous research [26,46,47]. An examination of the optimal molecular structure of the six coumarin derivatives revealed the existence of a partly double N–N bond, which facilitated the delocalization of π-electrons between the coumarin base and the fused aromatic ring. Based on the observed value of the dihedral angle C2–C3–C1′–N1, which is around ~130°, it is apparent that all the structures display deviations from planarity. Moreover, the determined N2–C7″–C1″–C6″ dihedral angles within the range of 152–179° for all the examined compounds suggest a deviation from planarity and a reduction in the electronic distribution within the compounds, which significantly affects their antioxidant activity. The obtained geometries were used for a deeper study of the mechanisms of the antioxidant action.

3.4.1. HAT Mechanism

The analysis of the BDE values for homolytic O–H cleavage showed that the reactivity decreased in the following order: 3d > 3c > 3f > 3b > 3a > 3e (Table 5). The reactivity of a compound is correlated with the stability of the newly formed radical species. Therefore, the NBO spin density of the studied radical species was analyzed. Figure S1 shows that the unpaired electron was mainly delocalized over the oxygen undergoing hydrogen abstraction and the ortho and para carbon atoms. Obviously, the better the delocalization of the unpaired electron, the more stable the obtained radical. The radicals formed from the o- and p-hydroxyl groups are the most stable because their unpaired electrons are delocalized over the benzene ring. The spin density values of the newly formed radicals increased in the following order: 3d-2O (0.284e) > 3f (0.306e) > 3c-4O (0.312e) > 3b (0.327e) > 3d-3O (0.330e) > 3c-3O (0.341e) > 3a (0.358e). The oxygen atom of the strongest 3d radical scavenger has the lowest spin density. The lower BDE value for 3d can be attributed to the stability of the radical species due to the free rotation of the C7″–C1″ bond, as well as the formation of an intramolecular hydrogen bond. Since the BDE values in benzene are comparatively lower than the PA and ETE values, it can be concluded that HAT is likely to be the dominant mechanism in nonpolar solvents.
On the other hand, the calculated BDE values for the homolytic cleavage of N–H bonds exhibited an increasing trend in the following order: 3e > 3b > 3d> 3f > 3c > 3a. The limited delocalization of the electron and the reduced stability of radicals arising from the homolytic cleavage of the N2–H bond can be attributed to the higher spin density values observed for N-centered radicals compared with O-centered radicals (Figure S2). This discrepancy is exemplified by the higher BDE values associated with N-centered radicals.
Generally, the BDE values obtained for compounds 3af exhibit similarities to the BDE values reported for the conventional antioxidants apigenin (372 and 368 kJ mol−1) and naringenin (368 and 362 kJ mol−1) [48]. The BDE values show a noticeably lower value in the non-polar solvent relative to the values in methanol, as shown in Table 5. These results are expected and in good agreement with the results of earlier research [26,36,37,38,39,49,50,51,52,53].

3.4.2. SET-PT Mechanism

The SET-PT mechanism is often thermodynamically least likely due to high IP values [44]. However, it is important to note that there exist certain instances when the SET-PT mechanism might exhibit dominance [54,55,56,57,58,59,60]. In methanol, the first step of the SET-PT mechanism for all investigated compounds, which is described by IP values, is extremely endothermic, whereas the second step is significantly exothermic in methanol (Table 5). As anticipated, the IP values were approximately 100 kJ mol−1 greater in the non-polar solvent than in methanol. Due to their high values, however, it can be inferred that the SET-PT mechanism is not a viable reaction pathway in either solvent. The theoretically predicted reactivity order was 3f > 3e > 3a, 3c > 3b > 3d. Although the studied compounds had lower and even exergonic PDE values, the initial step of the SET-PT pathway was the limiting factor for the antioxidant activity of these compounds via this mechanism.

3.4.3. SPLET Mechanism

The heterolytic cleavage of the O–H and N–H bonds of investigated compounds is the first step of the SPLET mechanism, which leads to the formation of the corresponding anions. The calculated thermodynamic parameters for the heterolytic cleavage of the O–H and N–H groups are shown in Table 5. The PA values of the O–H group were lower than those of the N–H group. The order of activity based on the PA value for the O–H group was 3d > 3b > 3c > 3a > 3f. The difference in reactivity was a consequence of the stabilization of the newly formed anionic species (Figures S3 and S4). According to the PA values in methanol (Table 5), the heterolytic cleavage of the O2′–H bond in 3d is favourable. The negative charge of this anion is distributed over O7″, O2″, O3″, and C5″. This anion is additionally stabilized with the hydrogen bonds. As expected, the stabilization of the anionic species in the polar solvent was much more pronounced, which is reflected in the lower PA values. Because the BDE values in benzene are lower than the PA and ETE values, it is plausible that HAT predominates in nonpolar solvents. The PA values in methanol were considerably lower than the BDE values; however, the ETE values were higher, but still comparable with the BDEs. Therefore, it is plausible to assume that HAT and SPLET compete within polar media.

3.5. Radical Scavenging Activity toward DPPH Radical

Potential free radical scavenging mechanisms are discussed based on the reaction free energies (ΔrG) following the HAT and SET-PT mechanisms (Figure 3). Calculations were performed using methanol as the solvent, mimicking the in vitro DPPH test conditions. The difference in energies between the reactants and products serves as the primary criterion for determining the probability of the occurrence of the reaction. The lower value of thermodynamic parameters describing HAT and SET mechanisms indicates the favorability of the pathway by which reaction occurs. The optimized geometries of the reactants involved in the reaction, namely DPPH, DPPH2, and DPPH are presented in Figure S5.
An examination of the results presented in Table 6 shows that all the investigated compounds exhibited positive ∆rGSET values (>132 kJ mol−1). This implies that the reactions involving the inactivation of the DPPH radical through the SET-PT mechanism are characterized by endergonic values. Based on the observed data, it can be concluded that the reactivity of the compounds decreased in the following order: 3f > 3e > 3a, 3c > 3b > 3d. The significant values of ∆rGSET clearly demonstrate that the SET-PT mechanism can be disregarded considering DPPH radical scavenging, in agreement with the thermodynamic data presented in Table 6.
ΔrGHAT values have been identified as a reliable measure for assessing the ability of coumarin derivatives to inactivate DPPH. The observed antioxidant activity, as assessed by in vitro measurements, exhibited the same trend as the ΔrGHAT values obtained in methanol: 3d > 3c > 3f > 3b > 3a > 3e. This finding further substantiates the notion that both this parameter and spin density possess the capacity to elucidate the reactivity with respect to DPPH. The enhanced antioxidant activity of compounds 3b and 3f has been previously highlighted because of the intramolecular stabilization of the radical species produced. Based on the aforementioned information, it is reasonable to anticipate that the DPPH present in methanol will undergo scavenging via the HAT mechanistic pathway.

3.6. Radical Scavenging Activity toward ABTS Radical Cation

ABTS•+ inactivation was evaluated using two different reaction pathways, namely HAT and SET-PT. Methanol was used as the solvent to replicate the experimental conditions (Figure 4). Table 7 lists the estimated values of the thermodynamic parameters. The optimized geometries of ABTS•+, ABTS+, and ABTS are shown in Figure S6.
The transfer of a hydrogen atom from the –OH groups of A–OH to ABTS•+, with the formation of A–O and ABTS+, is a thermodynamically favored reaction. The exergonic ΔrGHAT values of compound 3d (−13 kJ mol−1) demonstrated the highest efficiency in neutralizing radical species, which aligns with the findings from the experimental data.
On the other hand, electron transfer from A–OH to ABTS•+ with the formation of A-OH+ and ABTS+ thermodynamics was favored in all the investigated compounds. Based on an analysis of the ΔrGSET values, it can be concluded that compound 3f (−155 kJ mol−1) exhibits the highest electron transfer ability. However, endergonic ΔrGPT values (>118 kJ mol−1) represent a limiting factor for the manifestation of antioxidant activity through the SET-PT mechanism. Based on the sum ΔrGSET + ΔrGPT, it can be concluded that the reactivity of the compounds under investigation follows a decreasing trend in the sequence 3d > 3c > 3f > 3b > 3a > 3e, which aligns with the experimental results. In summary, it can be concluded that HAT and SET-PT are in competition when it comes to the inactivation of the ABTS•+ radical in methanol.

4. Conclusions

In this study, six coumarin N-acylhydrazone derivatives have been synthesized in high yields and characterized using NMR and IR spectroscopy, as well as elemental analysis. Also, the mechanism for the synthesis of coumarin N-acylhydrazone derivatives 3 in the presence of molecular iodine as a catalyst was proposed. All the obtained compounds were subjected to investigation of their antioxidant potential using DPPH, ABTS, and FRAP tests. The obtained in vitro results revealed that newly synthesized phenolic compounds 3c and 3d exerted excellent activities compared with reference compounds. The Gibbs free energies of reactions were used to determine the most likely mechanism of action. This methodology includes calculations of energy changes during a chemical reaction, providing insight into the feasibility of different reaction pathways. This study also highlighted the importance of considering solvent polarity and free radical species when studying antiradical properties. It was found that HAT and SPLET are viable molecular pathways for the radical scavenging activity of 3af in polar mediums, while the HAT is the dominant mechanism in non-polar environments. The changes in Gibbs free energies of reactions suggest the HAT and SET-PT as thermodynamically preferred mechanisms for ABTS•+ radical inactivation in methanol. Taking into account all the obtained results in terms of the synthesis and antioxidant potency of the isolated compounds, it can be concluded that significant progress has been achieved in relation to the previous study related to coumarin hydroxy benzohydrazides [26]. Namely, the reactions lasted for much shorter periods, the yields of the isolated compounds were very good, and the antioxidant potential of some of the analogues was better.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/antiox12101858/s1. Figure S1: NBO spin distribution for formed O-centered radical species; Figure S2: NBO spin distribution for formed N-centered radical species; Figure S3: NBO charge distribution for formed O-centered anionic species; Figure S4: NBO charge distribution for formed N-centered anionic species; Figure S5: Optimized geometry of radical, neutral and anionic DPPH species at M06-2X/6-311++G(d,p) level of theory in methanol (SMD solvation model); Figure S6: Optimized geometry of radical cation, neutral and cation ABTS species at M06-2X/6-311++G(d,p) level of theory in methanol (SMD solvation model); Table S1: DPPH scavenging activity of products 3c and 3d, as well as referent compounds at concentrations close to the IC50 value; Table S2: ABTS radical cation scavenging activity of products 3c, 3d, and referent compound Trolox at concentrations close to the IC50 value.

Author Contributions

Conceptualization, Z.S.M., D.A.M. and D.M.S.; methodology, Z.S.M., D.M.S. and E.H.A.; software, Ž.B.M., A.D.A. and M.R.A.; validation, Z.S.M. and Z.D.; formal analysis, E.H.A. and A.D.A.; investigation, D.M.S. and E.H.A.; resources, Z.S.M. and D.A.M.; writing—original draft preparation, D.M.S. and Ž.B.M.; writing—review and editing, Z.S.M., D.A.M. and E.H.A.; visualization, A.D.A. and M.R.A.; supervision, Z.S.M.; project administration, D.M.S. and D.A.M.; funding acquisition, Z.S.M. and D.A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Serbian Ministry of Science, Technological Development, and Innovations Agreements No. 451-03-47/2023-01/200378, 451-03-47/2023-01/200252, and the project of the Center for Scientific Research of SANU and the University of Kragujevac No. VI-03-232.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Murray, D.H.; Mendez, J.; Brown, S.A. The Natural Coumarins: Occurrence, Chemistry and Biochemistry; J. Wiley & Sons: New York, NY, USA, 1982; p. 1. [Google Scholar]
  2. O’Kennedy, R.; Thornes, R.D. Coumarins: Biology, Applications and Mode of Action; J. Wiley & Sons: Chichester, UK, 1997; p. 1. [Google Scholar]
  3. Choi, J.; Lee, K.-T.; Ka, H.; Jung, W.T.; Jung, H.-J.; Park, H.-J. Constituents of the Essential Oil of the Cinnamomum cassia Stem Bark and the Biological Properties. Arch. Pharm. Res. 2001, 24, 418–423. [Google Scholar] [CrossRef] [PubMed]
  4. Sharifi-Rad, J.; Cruz-Martins, N.; López-Jornet, P.; Pons-Fuster Lopez, E.; Harun, N.; Yeskaliyeva, B.; Beyatli, A.; Sytar, O.; Shaheen, S.; Sharopov, F.; et al. Natural Coumarins: Exploring the Pharmacological Complexity and Underlying Molecular Mechanisms. Oxidative Med. Cell. Longev. 2021, 2021, 6492346. [Google Scholar] [CrossRef] [PubMed]
  5. Bhatia, R.; Rawal, R.K. Coumarin hybrids: Promising scaffolds in the treatment of breast cancer. Mini-Rev. Med. Chem. 2019, 19, 1443–1458. [Google Scholar] [CrossRef]
  6. Al-Amiery, A.A.; Al-Majedy, Y.K.; Kadhum, A.A.H.; Mohamad, A.B. Novel macromolecules derived from coumarin: Synthesis and antioxidant activity. Sci. Rep. 2015, 5, 11825. [Google Scholar] [CrossRef] [PubMed]
  7. Senol, F.S.; Yilmaz, G.; Sener, B.; Koyuncu, M.; Orhan, I. Preliminary screening of acetylcholinesterase inhibitory and antioxidant activities of Anatolian Heptaptera species. Pharm. Biol. 2010, 48, 337–341. [Google Scholar] [CrossRef]
  8. Kancheva, V.D.; Boranova, P.V.; Nechev, J.T.; Manolov, I.I. Structure-activity relationships of new 4-hydroxy bis-coumarins as radical scavengers and chainbreaking antioxidants. Biochimie 2010, 92, 1138–1146. [Google Scholar] [CrossRef]
  9. Amin, K.M.; Taha, A.M.; George, R.F.; Mohamed, N.M.; Elsenduny, F.F. Synthesis, antitumor activity evaluation, and DNA-binding study of coumarin-based agents. Arch. Pharm. 2018, 351, 1700199. [Google Scholar] [CrossRef]
  10. Arya, A.K.; Rana, K.; Kumar, M. A facile synthesis and anticancer activity evaluation of spiro analogs of benzothiazolylchromeno/pyrano derivatives. Lett. Drug Des. Discov. 2014, 11, 594–600. [Google Scholar] [CrossRef]
  11. Ambekar, S.P.; Mohan, C.D.; Shirahatti, A.; Kumar, M.K.; Rangappa, S.; Mohan, S.; Basappa Kotresh, O.; Rangappa, K.S. Synthesis of coumarin-benzotriazole hybrids and evaluation of their anti-tubercular activity. Lett. Org. Chem. 2018, 15, 23–31. [Google Scholar] [CrossRef]
  12. Lin, P.Y.; Yeh, K.-S.; Su, C.-L.; Sheu, S.-Y.; Chen, T.; Ou, K.-L.; Lin, M.-H.; Lee, L.-W. Synthesis and antibacterial activities of novel 4-hydroxy-7-hydroxy- and 3-carboxycoumarin derivatives. Molecules 2012, 17, 10846–10863. [Google Scholar] [CrossRef]
  13. Chohan, Z.H.; Shaikh, A.U.; Rauf, A.; Supuran, C.T. Antibacterial, antifungal and cytotoxic properties of novel N-substituted sulfonamides from 4-hydroxycoumarin. J. Enzyme Inhib. Med. Chem. 2006, 21, 741–748. [Google Scholar] [CrossRef]
  14. Shi, X.; Lv, C.W.; Li, J.; Hou, Z.; Yang, X.H.; Zhang, Z.D.; Luo, X.X.; Yuan, Z.; Li, M.K. Synthesis, photoluminescent, antibacterial activities and theoretical studies of three novel coumarin and dihydropyran derivatives containing a triphenylamine group. Res. Chem. Intermediat. 2015, 41, 8965–8974. [Google Scholar] [CrossRef]
  15. Zhang, R.R.; Liu, J.; Zhang, Y.; Hou, M.Q.; Zhang, M.Z.; Zhou, F.; Zhang, W.H. Microwave-assisted synthesis and antifungal activity of novel coumarin derivatives: Pyrano [3,2-c] chromene-2, 5-diones. Eur. J. Med. Chem. 2016, 116, 76–83. [Google Scholar] [CrossRef] [PubMed]
  16. Shen, Y.F.; Liu, L.; Feng, C.Z.; Hu, Y.; Chen, C.; Wang, G.X.; Zhu, B. Synthesis and antiviral activity of a new coumarin derivative against spring viraemia of carp virus. Fish Shellfish Immun. 2018, 81, 57–66. [Google Scholar] [CrossRef] [PubMed]
  17. Wardrop, D.; Keeling, D. The story of the discovery of heparin and warfarin. Br. J. Haematol. 2008, 141, 757–763. [Google Scholar] [CrossRef]
  18. Holbrook, A.M.; Pereira, J.A.; Labiris, R.; McDonald, H.; Douketis, J.D.; Crowther, M.; Wells, P.S. Systematic overview of warfarin and its drug and food interactions. Arch. Intern. Med. 2005, 165, 1095–1106. [Google Scholar] [CrossRef]
  19. Abdelhafez, O.M.; Amin, K.M.; Batran, R.Z.; Maher, T.J.; Nada, S.A.; Sethumadhavan, S. Synthesis, Anticoagulant and PIVKA-II induced by new 4-hydroxycoumarin derivatives. Bioorg. Med. Chem. 2010, 18, 3371–3378. [Google Scholar] [CrossRef]
  20. O’Reilly, R.A.; Ohms, J.I.; Motley, C.H. Studies on coumarin anticoagulant drugs: Heat of interaction of sodium warfarin and human plasma albumin by heat burst microcalorimetry. J. Biol. Chem. 1969, 244, 1303–1305. [Google Scholar] [CrossRef]
  21. Braun, A.E.; Gonzalez, A.G. Coumarins. Nat. Prod. Rep. 1997, 14, 465–475. [Google Scholar] [CrossRef]
  22. Angelova, V.T.; Vassilev, N.G.; Nikolova-Mladenova, B.; Vitas, J.; Malbasa, R.; Momekov, G.; Djukic, M.; Saso, L. Antiproliferative and antioxidative effects of novel hydrazone derivatives bearing coumarin and chromene moiety. Med. Chem. Res. 2016, 25, 2082–2089. [Google Scholar]
  23. Taha, M.; Adnan Ali Shah, S.; Afifi, M.; Imran, S.; Sultan, S.; Rahim, F.; Ismail, N.H.; Khan, K.M. Synthesis, molecular docking study and thymidine phosphorylase inhibitory activity of 3-formylcoumarin derivatives. Bioorg. Chem. 2018, 78, 17–23. [Google Scholar] [CrossRef]
  24. Kotali, A.; Nasiopoulou, D.A.; Harris, P.A.; Helliwell, M.; Joule, J.A. Transformation of a hydroxyl into an acyl group on α-pyrone ring: A novel route to 3,4-diacylcoumarins. Tetrahedron 2012, 68, 761–766. [Google Scholar] [CrossRef]
  25. Kotali, A.; Nasiopoulou, D.A.; Tsoleridis, C.A.; Harris, P.A.; Kontogiorgis, C.A.; Hadjipavlou-Litina, D.J. Antioxidant Activity of 3-[N-(Acylhydrazono) ethyl]-4-hydroxy-coumarins. Molecules 2016, 21, 138. [Google Scholar] [CrossRef] [PubMed]
  26. Antonijević, M.R.; Simijonović, D.M.; Avdović, E.H.; Ćirić, A.; Petrović, Z.D.; Marković, J.D.; Stepanić, V.; Marković, Z.S. Green One–Pot Synthesis of Coumarin-Hydroxybenzohydrazide Hybrids and Their Antioxidant Potency. Antioxidants 2021, 10, 1106. [Google Scholar] [CrossRef]
  27. Pangal, A.; Shaikh, J.A.; Muiz, G.V.; Mane Ahmed, K. Novel 3-acetylcoumarin Schiff’s base synthesis from different acid hydrazide. Int. Res. J. Pharm. 2013, 4, 108–110. [Google Scholar] [CrossRef]
  28. Prihantini, A.I.; Tachibana, S.; Itoh, K. Antioxidant active compounds from elaeocarpussylvestris and their relationship between structure and activity. Procedia Environ. Sci. 2015, 28, 758–768. [Google Scholar] [CrossRef]
  29. Pontiki, E.; Hadjipavlou-Litina, D.; Litinas, K.; Geromichalos, G. Novel Cinnamic Acid Derivatives as Antioxidant and Anticancer Agents: Design, Synthesis and Modeling Studies. Molecules 2014, 19, 9655–9674. [Google Scholar] [CrossRef]
  30. Pownall, T.L.; Udenigwe, C.C.; Aluko, R.E. Amino acid composition and antioxidant properties of pea seed (Pisum sativum L.) enzymatic protein hydrolysate fractions. J. Agric. Food Chem. 2010, 58, 4712–4718. [Google Scholar] [CrossRef]
  31. Marc, G.; Stana, A.; Franchini, A.H.; Vodnar, D.C.; Barta, G.; Terti, M.; Santa, I.; Cristea, C.; Pîrnau, A.; Ciorîta, A.; et al. Phenolic Thiazoles with Antioxidant and Antiradical Activity. Synthesis, In Vitro Evaluation, Toxicity, Electrochemical Behavior, Quantum Studies and Antimicrobial Screening. Antioxidants 2021, 10, 1707. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.W.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  33. Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007. [Google Scholar] [CrossRef]
  34. Becke, A.D.; Johnson, E.R. A density-functional model of the dispersion interaction. J. Chem. Phys. 2005, 123, 154101. [Google Scholar] [CrossRef] [PubMed]
  35. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  36. Bernales, V.S.; Marenich, A.V.; Contreras, R.; Cramer, C.J.; Truhlar, D.G. Quantum mechanical continuum solvation models for ionic liquids. J. Phys. Chem. B 2012, 116, 9122–9129. [Google Scholar] [CrossRef]
  37. Milanović, Ž.; Tošović, J.; Marković, S.; Marković, Z. Comparison of the scavenging capacities of phloroglucinol and 2, 4, 6-trihydroxypyridine towards HO radical: A computational study. RSC Adv. 2020, 10, 43262–43272. [Google Scholar] [CrossRef] [PubMed]
  38. Antonijević, M.R.; Avdović, E.H.; Simijonović, D.M.; Milanović, Ž.B.; Amić, A.D.; Marković, Z.S. Radical scavenging activity and pharmacokinetic properties of coumarin–hydroxybenzohydrazide hybrids. Int. J. Mol. Sci. 2022, 23, 490. [Google Scholar] [CrossRef] [PubMed]
  39. Milanović, Ž.; Dimić, D.; Žižić, M.; Milenković, D.; Marković, Z.; Avdović, E. Mechanism of antiradical activity of newly synthesized 4, 7-dihydroxycoumarin derivatives-experimental and kinetic DFT study. Int. J. Mol. Sci. 2021, 22, 13273. [Google Scholar] [CrossRef]
  40. Milanović, Ž.; Dimić, D.; Antonijević, M.; Žižić, M.; Milenković, D.; Avdović, E.; Marković, Z. Influence of acid-base equilibria on the rate of the chemical reaction in the Advanced Oxidation Processes: Coumarin derivatives and hydroxyl radical. Chem. Eng. J. 2023, 453, 139648. [Google Scholar] [CrossRef]
  41. Wang, G.; Xue, Y.; An, L.; Zheng, Y.; Dou, Y.; Zhang, L.; Liu, Y. Theoretical study on the structural and antioxidant properties of some recently synthesised 2,4,5-trimethoxy chalcones. Food Chem. 2015, 171, 89–97. [Google Scholar] [CrossRef]
  42. Wang, L.; Yang, F.; Zhao, X.; Li, Y. Effects of nitro- and amino-group on the antioxidant activity of genistein: A theoretical study. Food Chem. 2019, 275, 339–345. [Google Scholar] [CrossRef]
  43. Marković, Z.; Tošović, J.; Milenković, D.; Marković, S. Revisiting the solvation enthalpies and free energies of the proton and electron in various solvents. Comp. Theor. Chem. 2016, 1077, 11–17. [Google Scholar] [CrossRef]
  44. Glendening, E.D.; Landis, C.R.; Weinhold, F. NBO 6.0: Natural bond orbital analysis program. J. Comp. Chem. 2013, 34, 1429–1437. [Google Scholar] [CrossRef] [PubMed]
  45. Dimić, D.; Milenković, D.; Marković, J.D.; Marković, Z. Antiradical activity of catecholamines and metabolites of dopamine: Theoretical and experimental study. Phys. Chem. Chem. Phys. 2017, 19, 12970–12980. [Google Scholar] [CrossRef] [PubMed]
  46. Milanović, Ž.B.; Dimić, D.S.; Avdović, E.H.; Milenković, D.A.; Dimitrić Marković, J.; Klisurić, O.R.; Trifunović, S.R.; Marković, Z.S. Synthesis and comprehensive spectroscopic (X-ray, NMR, FTIR, UV–Vis), quantum chemical and molecular docking investigation of 3-acetyl-4-hydroxy-2-oxo-2H-chromen-7-yl acetate. J. Mol. Struct. 2021, 1225, 129256. [Google Scholar] [CrossRef]
  47. Avdović, E.H.; Milanović, Ž.B.; Molčanov, K.; Roca, S.; Vikić-Topić, D.; Mrkalić, E.M.; Jelić, R.M.; Marković, Z.S. Synthesis, characterization and investigating the binding mechanism of novel coumarin derivatives with human serum albumin: Spectroscopic and computational approach. J. Molec. Struct. 2022, 1254, 132366. [Google Scholar] [CrossRef]
  48. Zheng, Y.Z.; Dengb, G.; Chena, D.F.; Guoa, R.; Laia, R.C. The influence of C2=C3 double bond on the antiradical activity of flavonoid: Different mechanisms analysis. Phytochemistry 2019, 157, 1–7. [Google Scholar] [CrossRef] [PubMed]
  49. Bakhouche, K.; Dhaouadi, Z.; Jaidane, N.; Hammoutène, D. Comparative antioxidant potency and solvent polarity effects on HAT mechanisms of tocopherols. Comput. Theor. Chem. 2015, 1060, 58–65. [Google Scholar] [CrossRef]
  50. Salamone, M.; Bietti, M. Reaction pathways of alkoxyl radicals. The role of solvent effects on C–C bond fragmentation and hydrogen atom transfer reactions. Synlett 2014, 25, 1803–1816. [Google Scholar]
  51. Litwinienko, G.; Ingold, K.U. Abnormal solvent effects on hydrogen atom abstraction. 2. Resolution of the curcumin antioxidant controversy. The role of sequential proton loss electron transfer. J. Org. Chem. 2004, 69, 5888–5896. [Google Scholar] [CrossRef]
  52. MacLean, P.D.; Chapman, E.E.; Dobrowolski, S.L.; Thompson, A.; Barclay, L.R.C. Pyrroles as antioxidants: Solvent effects and the nature of the attacking radical on antioxidant activities and mechanisms of pyrroles, dipyrrinones, and bile pigments. J. Org. Chem. 2008, 73, 6623–6635. [Google Scholar] [CrossRef]
  53. Shang, Y.J.; Qian, Y.P.; Liu, X.D.; Dai, F.; Shang, X.L.; Jia, W.Q.; Liu, Q.; Fang, J.G.; Zhou, B. Radical-scavenging activity and mechanism of resveratrol-oriented analogues: Influence of the solvent, radical, and substitution. J. Org. Chem. 2009, 74, 5025–5031. [Google Scholar] [CrossRef]
  54. Hamlaoui, I.; Bencheraiet, R.; Bensegueni, R.; Bencharif, M. Experimental and theoretical study on DPPH radical scavenging mechanism of some chalcone quinoline derivatives. J. Mol. Struct. 2018, 1156, 385–389. [Google Scholar] [CrossRef]
  55. Rimarčík, J.; Lukeš, V.; Klein, E.; Ilčin, M. Study of the solvent effect on the enthalpies of homolytic and heterolytic N–H bond cleavage in p-phenylenediamine and tetracyano-p-phenylenediamine. J. Mol. Struct. Theochem. 2010, 952, 25–30. [Google Scholar] [CrossRef]
  56. Chen, J.; Yang, J.; Ma, L.; Li, J.; Shahzad, N.; Kim, C.K. Structure-antioxidant activity relationship of methoxy, phenolic hydroxyl, and carboxylic acid groups of phenolic acids. Sci. Rep. 2020, 10, 2611. [Google Scholar] [CrossRef] [PubMed]
  57. Klein, E.; Lukeš, V. DFT/B3LYP study of the substituent effect on the reaction enthalpies of the individual steps of single electron transfer-proton transfer and sequential proton loss electron transfer mechanisms of phenols antioxidant action. J. Phys. Chem. A 2006, 110, 12312–12320. [Google Scholar] [CrossRef] [PubMed]
  58. Zheng, Y.Z.; Deng, G.; Liang, Q.; Chen, D.F.; Guo, R.; Lai, R.C. Antioxidant activity of quercetin and its glucosides from propolis: A theoretical study. Sci. Rep. 2017, 7, 7543. [Google Scholar] [CrossRef] [PubMed]
  59. Alisi, I.O.; Uzairu, A.; Abechi, S.E. Free radical scavenging mechanism of 1,3,4-oxadiazole derivatives: Thermodynamics of O–H and N–H bond cleavage. Heliyon 2020, 6, e03683. [Google Scholar] [CrossRef] [PubMed]
  60. Mittal, A.; Vashistha, V.K.; Das, D.K. Recent advances in the antioxidant activity and mechanisms of chalcone derivatives: A computational review. Free Radic. Res. 2022, 56, 378–397. [Google Scholar] [CrossRef]
Figure 1. Structures of some coumarin derivatives with clinical application.
Figure 1. Structures of some coumarin derivatives with clinical application.
Antioxidants 12 01858 g001
Scheme 1. Synthesis of coumarin N-acylhydrazone derivatives: 3-acetylcoumarin (1), benzoyl hydrazide (2), iodine as catalyst (10 mol%), absolute ethanol, and room temperature (RT).
Scheme 1. Synthesis of coumarin N-acylhydrazone derivatives: 3-acetylcoumarin (1), benzoyl hydrazide (2), iodine as catalyst (10 mol%), absolute ethanol, and room temperature (RT).
Antioxidants 12 01858 sch001
Scheme 2. The suggested mechanism for formation of coumarin N-acylhydrazone derivatives.
Scheme 2. The suggested mechanism for formation of coumarin N-acylhydrazone derivatives.
Antioxidants 12 01858 sch002
Figure 2. Optimized geometries of investigated compounds, 3a3f at M06-2X/6-311++G(d,p) level of theory in methanol (SMD solvation model). Legend: gray—carbon atom, white—hydrogen atom, red—oxygen atom, blue—nitrogen atom.
Figure 2. Optimized geometries of investigated compounds, 3a3f at M06-2X/6-311++G(d,p) level of theory in methanol (SMD solvation model). Legend: gray—carbon atom, white—hydrogen atom, red—oxygen atom, blue—nitrogen atom.
Antioxidants 12 01858 g002
Figure 3. The possible reaction mechanism: HAT and SET-PT, between DPPH radical and investigation compounds 3a3f (A–OH).
Figure 3. The possible reaction mechanism: HAT and SET-PT, between DPPH radical and investigation compounds 3a3f (A–OH).
Antioxidants 12 01858 g003
Figure 4. The possible reaction mechanism: HAT and SET-PT, between ABTS radical cation and investigated compounds 3a3f (A–OH).
Figure 4. The possible reaction mechanism: HAT and SET-PT, between ABTS radical cation and investigated compounds 3a3f (A–OH).
Antioxidants 12 01858 g004
Table 1. Optimization of reaction conditions.
Table 1. Optimization of reaction conditions.
EntryCatalyst (mol%)Yield (%)
1Nonetrace a,c/65 b,c
2CH3COOH (10 mol%)30 a,c/75 b,c
3CH3COOH (20 mol%)42 a,c/80 b,c
4I2 (5 mol%)85 a,d
5I2 (10 mol%)91 a,d
6I2 (20 mol%)91 a,d
Reaction conditions: 1 (1 mmol), 2a (1 mmol), 3 mL EtOH; a Reaction performed at the room temperature; b Reaction performed by heating under reflux; c Reaction finished for 10 h; d Reaction finished for 20 min.
Table 2. Isolated yields of coumarin N-acylhydrazone derivatives 3.
Table 2. Isolated yields of coumarin N-acylhydrazone derivatives 3.
CompoundTime (Min)Yield (%)
3a2091
3b2093
3c2082
3d3080
3e3095
3f2083
Table 3. DPPH scavenging activity of products 3 and referent compounds.
Table 3. DPPH scavenging activity of products 3 and referent compounds.
CompoundDPPH Scavenging Ability (%)IC50 (µM)SF
25 µM50 µM100 µM
20 min60 min20 min60 min20 min60 min
3a1.0 ± 0.71.2 ± 0.71.3 ± 1.14.2 ± 1.11.6 ± 1.34.6 ± 1.8--
3b3.5 ± 2.33.8 ± 3.04.2 ± 0.98.6 ± 1.25.5 ± 0.98.6 ± 0.9--
3c94.2 ± 0.494.7 ± 0.994.1 ± 2.194.5 ± 1.898.4 ± 1.298.6 ± 1.12.1 ± 0.16.0
3d97.9 ± 1.798.1 ± 0.698.3 ± 1.498.6 ± 0.898.4 ± 0.398.7 ± 0.43.2 ± 0.13.9
3e0.5 ± 0.47.4 ± 0.31.08 ± 0.310.8 ± 1.51.4 ± 0.435.2 ± 0.1--
3f1.6 ± 0.129.2 ± 0.33.4 ± 0.329.7 ± 2.38.7 ± 2.531.8 ± 1.7--
NDGA94.6 ± 0.794.6 ± 0.694.2 ± 0.794.2 ± 0.794.5 ± 0.294.1 ± 0.71.7 ± 0.17.4
Quercetin95.3 ± 0.895.1 ± 0.996.8 ± 1.096.5 ± 0.995.1 ± 0.995.4 ± 0.81.9 ± 0.16.6
Table 4. ABTS radical cation scavenging activity and ferric ion reducing capacity of products 3 and corresponding referent compounds.
Table 4. ABTS radical cation scavenging activity and ferric ion reducing capacity of products 3 and corresponding referent compounds.
CompoundABTS Radical Cation Scavenging ActivityFerric Ion Reducing Capacity
Scavenging Ability (%)IC50 (µM)
25 µM50 µM100 µM% AAa
3a9.7 ± 2.513.1 ± 2.320.5 ± 2.7-137.13
3b15.3 ± 0.918.5 ± 1.231.1 ± 1.8-128.96
3c97.8 ± 1.098.1 ± 1.299.2 ± 1.02.2 ± 0.2718.41
3d98.4 ± 2.198.7 ± 0.798.8 ± 1.43.0 ± 0.1510.96
3e8.5 ± 1.216.5 ± 1.225.4 ± 0.8-153.77
3f42.4 ± 1.143.2 ± 0.453.6 ± 2.0-216.34
Trolox97.4 ± 0.299.3 ± 0.199.5 ± 0.31.3 ± 0.1-
Table 5. Calculated thermodynamic parameters (in kJ mol−1) of antioxidant mechanisms of investigated compounds 3af. DFT calculations were performed at SMD/M06-2X/6-311++G(d,p) level of theory in methanol and benzene as solvents.
Table 5. Calculated thermodynamic parameters (in kJ mol−1) of antioxidant mechanisms of investigated compounds 3af. DFT calculations were performed at SMD/M06-2X/6-311++G(d,p) level of theory in methanol and benzene as solvents.
CompoundsPositionMethanolBenzene
BDEIPPDEPAETEBDEIPPDEPAETE
3a4″–OH38652430143411375646143411359
N2–H40346167404383150429367
3b2″–OH37853016124422377659131389400
N2–H40038149419384138400397
3c3″–OH3575241139385346646113402357
4″–OH3571131394345112389369
N2–H40347167404385152427370
3d2″–OH347532−1711240334466592369389
3″–OH360−4154373369117455327
N2–H40137146423387135394406
3eN2–H39852245158408385642156430368
3f4″–OH36550923144349369664118443349
N2–H40261167364383132432364
Table 6. Calculated thermodynamic parameters (in kJ mol−1) of antioxidant mechanisms between investigated compounds 3af and 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical at SMD/M06-2X/6-311++G(d,p) level of theory in methanol.
Table 6. Calculated thermodynamic parameters (in kJ mol−1) of antioxidant mechanisms between investigated compounds 3af and 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical at SMD/M06-2X/6-311++G(d,p) level of theory in methanol.
CompoundsPositionMechanisms
HATSETPT
ΔrGHATΔrGSETΔrGPT
3a4″–OH52147−95
N2–H69−78
3b2″–OH44153−109
N2–H67−86
3c3″–OH23147−124
4″–OH24−123
N2–H69−77
3d2″–OH14155−142
3″–OH26−129
N2–H68−87
3eN2–H65145−80
3f4″–OH31132−101
N2–H69−64
Table 7. Calculated thermodynamic parameters (in kJ mol−1) of antioxidant mechanisms between investigated compounds 3af and 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid) radical cation (ABTS•+) at SMD/M06-2X/6-311++G(d,p) level of theory in methanol.
Table 7. Calculated thermodynamic parameters (in kJ mol−1) of antioxidant mechanisms between investigated compounds 3af and 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid) radical cation (ABTS•+) at SMD/M06-2X/6-311++G(d,p) level of theory in methanol.
CompoundsPositionMechanisms
HATSETPT
ΔrGHATΔrGSETΔrGPT
3a4″–OH25−140165
N2–H42182
3b2″–OH17−134151
N2–H40173
3c3″–OH−4−140136
4″–OH−3137
N2–H42182
3d2″–OH−13−132118
3″–OH−1131
N2–H41172
3eN2–H38−142180
3f4″–OH4−155159
N2–H41196
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Simijonović, D.M.; Milenković, D.A.; Avdović, E.H.; Milanović, Ž.B.; Antonijević, M.R.; Amić, A.D.; Dolićanin, Z.; Marković, Z.S. Coumarin N-Acylhydrazone Derivatives: Green Synthesis and Antioxidant Potential—Experimental and Theoretical Study. Antioxidants 2023, 12, 1858. https://doi.org/10.3390/antiox12101858

AMA Style

Simijonović DM, Milenković DA, Avdović EH, Milanović ŽB, Antonijević MR, Amić AD, Dolićanin Z, Marković ZS. Coumarin N-Acylhydrazone Derivatives: Green Synthesis and Antioxidant Potential—Experimental and Theoretical Study. Antioxidants. 2023; 12(10):1858. https://doi.org/10.3390/antiox12101858

Chicago/Turabian Style

Simijonović, Dušica M., Dejan A. Milenković, Edina H. Avdović, Žiko B. Milanović, Marko R. Antonijević, Ana D. Amić, Zana Dolićanin, and Zoran S. Marković. 2023. "Coumarin N-Acylhydrazone Derivatives: Green Synthesis and Antioxidant Potential—Experimental and Theoretical Study" Antioxidants 12, no. 10: 1858. https://doi.org/10.3390/antiox12101858

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