Skip to main content
ACS Omega logoLink to ACS Omega
. 2021 Nov 17;6(47):31993–32004. doi: 10.1021/acsomega.1c04772

Investigation of 4-Hydrazinobenzoic Acid Derivatives for Their Antioxidant Activity: In Vitro Screening and DFT Study

Hatem A Abuelizz , Hanan A A Taie , Ahmed H Bakheit , Gamal A E Mostafa , Mohamed Marzouk §, Harunor Rashid , Rashad Al-Salahi †,*
PMCID: PMC8638017  PMID: 34870022

Abstract

graphic file with name ao1c04772_0011.jpg

Hydrazinobenzoic acid derivatives with isothiocyanate, benzylidene, and acid anhydride core units (113) were previously synthesized and fully characterized. Targets 113 were investigated for their antioxidant activities using different in vitro assays such as 1,1-diphenyl-1-picrylhydrazyl (DPPH) radical scavenging, 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS), ferric reducing antioxidant power (FRAP), and reducing power capability. All derivatives showed antioxidant properties in relation to the standard butylated hydroxylanisole (BHA). Superior antioxidant activities was observed for compounds 3 and 5–9 at a concentration of 20 μg/mL (70–72%) when tested by the DPPH method in comparison to BHA (92%), and compounds 1–10 showed the highest free radical quenching activity (80–85%) when examined by ABTS at 20 μg/mL in relation to BHA (85%). Density function theory (DFT) studies were carried out using the B3LYP/6-311G(d,p) level of theory. Several antioxidant descriptors were calculated for targets 113 compared with BHA. Targets 113 were proposed to exhibit their antioxidant activities via the following three proposed antioxidant mechanisms: single electron transfer (SET), hydrogen atom transfer (HAT), and sequential proton loss electron transfer (SPLET). The highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energies and electron levels for 113 were also determined.

1. Introduction

The development of several organic heterocyclic compounds, owing to their diverse biological properties, has gained much attention. Among them, hydrazine-, hydrazone-, and isothiocyanate-based compounds are found to have various applications in organic synthesis, medicinal chemistry, combinatorial chemistry, and supramolecular chemistry.13 Concerning organic synthesis, such compounds were proved to participate in the construction of different heterocyclic scaffold classes such as azetidines, coumarins, oxadiazolines, diazoles, triazoles, quinazolines, thiazolidines benzothiazines, and triazoloquinazolines.110 Their distinct biological significance was explained on the basis of the specific structural combination for a (−NH–N=CH−) or (N–N) group with a carbonyl or thiocarbonyl group and their reactivities toward both electrophiles and nucleophiles. Therefore, they are present in many bioactive compounds and demonstrate a wide variety of biological activities including anticancer, antioxidant, antimicrobial, antifungal, anti-inflammatory, anticonvulsant, antiviral, and antiprotozoal effects.916 The process of discovering antioxidants has been gaining growing interest due to their valuable protective roles in both pharmaceutical and food industries. In human physiology, they play an essential role against oxidative deterioration and oxidative stress-mediated mechanisms, whereas in the food system, antioxidants retard lipid peroxidation, prevent the formation of secondary lipid peroxidation products, maintain the flavor, color, and texture of the food product during storage, and reduce the interaction of lipid-derived carbonyls with proteins that is accompanied by alteration in protein function.17,18 There has been renewed interest to explore new antioxidant molecules either synthetically or naturally, so as to utilize them as the main agents to counter toxicity mediated by radicals.

Further evidence suggests that age-related human illnesses like heart disease, cancer, neurodegenerative, arthritis, immunosuppression, cerebral dysfunction, cataracts, and inflammation are the consequences of cellular damage by free radicals.19 Based on the aforementioned facts and our interest to discover novel synthetic antioxidants capable of preventing free radical-mediated damage and toxicity, we embarked on this experiment to evaluate the previously synthesized 4-hydrazinobenzoic acid derivatives 113 for their antioxidant activities using different in vitro assays such as DPPH, FRAP, ABTS, and reducing power capability. Moreover, the electronic and structural features of the investigated 113 have been evaluated by DFT.20 The bond dissociation enthalpy (BDE), proton dissociation enthalpy (PDE), ionization potential (IP), and proton affinity (PA) descriptors have been computed at the B3LYP/6-311G(d,p) level of theory to explain antioxidant mechanisms. For more clarification, the calculated reactivity descriptors such as dipole moment, electron affinity (EA), softness (S), hardness (η), electrophilicity (ω), electronegativity (χ), and chemical potential (μ) have been taken into account for targets 113 and also for BHA. In addition, the HOMO and LUMO energies were calculated.

2. Results and Discussion

2.1. DPPH Radical Scavenging Activity

In the published literature, several compounds incorporating hydrazine, hydrazone, and isothiocyanate units have been described as potential antioxidants. For instance, hydrazone heterocycles bearing an imidazole or benzimidazole entity showed the highest DPPH radical scavenging (IC50 = 22.4 ± 3.2 μM) and comparable ABTS scavenging (IC50 = 7.3 ± 0.8 μM) in relation to the reference BHT (70.8 ± 6.6 and 7.2 ± 1.7 μM) as a standard substance.16 Moreover, in comparison with quercetin as a reference drug, veratric acid (3,4-dimethoxybenzoic acid) derivatives incorporating benzylidene-hydrazine moieties were reported as promising tyrosinase inhibitors and potent free radical scavengers, giving an EC50 value of 0.0097 ± 0.0011 mM.21 In addition, triazoloquinazoline compounds resulting from a combination of hydrazinobenzoic acid with cyanoimdiocarbaonte demonstrated the highest capacity to deplete DPPH and free radicals (85–91%), in relation to BHT (93%), as a synthetic antioxidant substance.22 Introduction of the alkyl(aralkyl)isothiocyanate unit to the naphthaoic or anthranilic acid scaffold afforded promising antioxidant benzoquinazoline or quinazoline derivatives, which could be considered for developing useful templates for novel potent antioxidant compounds in the future.2326 The grafting degree of hydrazine-thiosemicarbazone groups in the 2-[phenylhydrazine (or hydrazine) thiosemicarbazone]-chitosan complex improved their free radical scavenging activity compared to chitosan, and their polymerization on chitosan increased their antioxidant activity obviously with an IC50 range of 0.263–0.531 mg/mL in comparison to ascorbic acid.19

Targets 113 (Scheme 1) were synthesized and characterized in our previous work.27

Scheme 1. Synthesis Routes for Targets 113.

Scheme 1

The DPPH steady radical could be attributed to the transfer of hydrogen or an electron to a stable compound. The DPPH methanolic solution demonstrated strong absorbance at 517 nm due to single electrons (odd electrons) or pairing electrons.2830 When the solution was mixed with suitable reducing agents (antioxidants), the color disappeared. Reduction in color is stoichiometrically increased with the number of electrons, and the decrease in the absorbance measurement could be directly compared with BHA. The DPPH radical scavenging ability of hydrazinobenzoic acid derivatives 113 was evaluated and compared with BHA as a standard antioxidant. A graphical representation of the antioxidant activity in parameters of % inhibition versus the concentration of compounds 113 is shown in Figure 1. The results indicate that most of the products possess high radical scavenging activity comparable to BHA (Figure 1). It is worth mentioning that compounds 3 and 59 exhibited a potent DPPH radical scavenging activity at a lower concentration (20 μg/mL) in relation to the other substituted compounds and the standard BHA at a concentration of 40 μg/mL. Their radical scavenging activities varied from 50.27 ± 0.15 to 72.29 ± 0.24%. Parent compound 1 exhibited a moderate DPPH radical scavenging ability of 41.48 ± 0.23%, while 11 and 12 showed the lowest abilities of 17.21 ± 0.16 and 18.52 ± 0.14%, respectively. Various substitutions (isothiocyanate and benzylidene) on the structure of 1 resulted in an increase in the radical scavenging activity, particularly for compounds 3 and 59 that showed intrinsically the highest activity at a very low concentration (20 μg/mL).

Figure 1.

Figure 1

DPPH radical scavenging activity of targets 113; data are mean ± standard deviation of triplicate experiments.

2.2. ABTS Radical Quenching Activity

Further examination of the antioxidant potency of targets 113 was performed by evaluation of their ABTS free radical scavenging ability, which is considered a commonly in vitro used method as well. The results presented in Figure 2 support the DPPH scavenging activity illustrated in Figure 1 and show that derivatives 710 possessed more superior scavenging capabilities compared to BHA activity. Except for compounds 1113, data illustrated in Figure 2 clearly indicate that all substituted derivatives showed an excellent ABTS radical scavenging activity. The hydroxyl-methoxy derivative (3) exhibited the highest ABTS radical quenching activity (85.19 ± 0.17%), mostly equal to that of BHA (85.20 ± 0.33) at the same concentration (20 μg/mL). Compounds 2 and 510 revealed potent ABTS radical scavenging ability in relation to BHA. Their radical scavenging effects ranged from 80.60 ± 0.15 for compound 6 to 84.34 ± 0.10% for 7 at the same concentration. Compound 1 demonstrated good antioxidant activity (74.52 ± 0.11%), while the substitution in compounds 11 and 13 resulted in moderate activities of 34.77 ± 0.15 and 46.55 ± 0.16%, respectively. The lowest ABTS radical scavenging ability was recorded by derivative 12 (18.51 ± 0.15%).

Figure 2.

Figure 2

ABTS radical scavenging ability of 1–13; data are mean ± standard deviation of triplicate experiments.

The DPPH scavenging effect of tested 113 was found to be in the descending order of BHA, 10, 9, 8, 7, 3, 2, 5, 6, 1, 4. On the other hand, the percentage values of ABTS scavenging activities of the same group were ordered in the descending manner of BHA, 3, 7, 8, 9, 10, 2, 5, 6, 1, 4. In our previous work, we reported that the radical scavenging activity of triazoloquinazolines22 depended mostly on the hydrazine and isothiocyanate units, which exerted their abilities as a better hydrogen supply or electron-donating group. The present outcomes affirmed the previous findings that 510 were the most active. The presence of hydrazine and isothiocyanate functional groups may result in maximum scavenging capabilities via their rules to stabilize the free radical form after electron donation or hydrogen supply.

2.3. FRAP

Iron plays a vital role in various normal neurological functions of living organisms; however, it can contribute to electron transfer reactions (redox reaction) with several substrates, making it an essential inducer for reactive species in humans.31 In view of this fact, the FRAP method was employed to estimate the capability of the electron donation of hydrazinobenzoic acid derivatives 113. The FRAP of a compound is a valuable indicator of its significant antioxidant ability. As shown in Figure 3, the majority of the examined substituents on the compound 1 structure led to a positive response to ferric reducing power capacity. Compound 1 showed a ferric reducing power capacity of 2864 ± 32.56 μmol Trolox/100 g, which is sharply increased to 4120 ± 20.53, 4080 ± 14.57, 4075 ± 11.06, and 4059 ± 33.42 μmol Trolox/100 g for compounds 7, 2, 3, and 5, respectively. An appropriate increase in FRAP by the substitution occurred in compounds 6, 8, 10, and 9 to achieve 3810 ± 10.50, 3733 ± 17.00, 3345 ± 20.52, and 3055 ± 21.01 μmol Trolox/100 g, respectively, at the same concentration (40 μg/mL). On the contrary, the substitutions in compounds 11, 12, and 13 caused a pronounced reduction in the ferric reducing power ability to 1047 ± 8.62, 946 ± 15.52, and 1263 ± 11.60 μmol Trolox/100 g, respectively, at 40 μg/mL. The FRAP of the targets was found in the order of 7 > 2 > 3 > 5 > 6 > 8 > 10 > 9 > 11 > 12 > 13.

Figure 3.

Figure 3

FRAP activity of targets 1–13; data are mean ± standard deviation of triplicate experiments.

2.4. Reducing Power Ability

In this method, the change of color from yellow to various green and blue shades relies on the reducing power of each substance at a concentration of 40 μg/mL. The antioxidant effect of 1–13 causes reduction of ferric (Fe3+) in the ferricyanide complex into the ferrous (Fe2+) form. Therefore, measuring the formation of Perls Prussian blue at 700 nm can control the concentration of Fe2+.31 The reducing power ability of targets 1–13 varied in high-to-moderate activity in comparison with BHA (Figure 4). Compound 1 demonstrated a high reducing power capability of 1.06 ± 0.03 at an absorbance of 700 nm. Substitutions on compound 1 resulted in a slight increase in the reducing power capability as observed by compounds 3, 2, and 5 (1.23 ± 0.04, 1.17 ± 0.02, and 1.10 ± 0.04, respectively), relative to BHA (1.35 ± 0.06) at the same concentration (40 μg/mL). Compounds 610 exhibited moderate reducing power activity ranging from 0.86 ± 0.01 to 1.02 ± 0.02. On the other hand, compounds 11, 12, and 13 recorded the lowest reducing power ability among all investigated 4-hydrazinobenzoic acid derivatives and showed the lowest antioxidant activity in all of the abovementioned assays.

Figure 4.

Figure 4

Reducing power ability of targets 1–13; data are mean ± standard deviation of triplicate experiments.

The strong correlations between the antioxidant ability assayed by ABTS, DPPH, FRAP, and reducing power capacity methods and compounds 113 (Figures 14) indicate that the hydrazine, isothiocyanate, and hydrazone units largely contribute to the antioxidant activities of 113. The presence of an anhydride moiety in 1113 does not offer advantages in the antioxidant profile. The results are in accordance with several studies, indicating that hydrazine and isothiocyanate core units are major antioxidant constituents in heterocyclic compounds.

2.5. DFT Study

2.5.1. Geometry Optimization and Reactivity Descriptors

The physicochemical descriptors like BDE, PDE, PA, IP, ETE, η, S, and μ (Table S1) were employed to visualize the existing relation between 1–13 structures and chemical reactivity. Accurate data for structural and electronic features of antioxidants are of great significance in describing their scavenging behavior. Thus, various initial geometries of the target compounds were selected to undergo full optimization using the B3LYP/6-311G(d,p) level of theory. In the present work, full optimization of compounds 1, 24, 510, and 11–13 was considered (Figure S1), and the orientation possibilities as H1 (aniline–Ar–NH), H2 (hydrazine–N–NH or hydrazinecarbothioamide N–NH–CS–NH), and H3 (phenol Ar–OH) were chosen.

Furthermore, in the presence of a weak external electric field, the electronic cloud system can be distorted from its normal shape.32 Different factors like compound complexity, molecular size, and the number of electrons can influence the polarizability parameter. The mean polarizability and dipole moment values are listed in Table 2. The tendency of charge distribution is accompanied by molecule polarizability. In our study, most of the target compounds showed higher polarizability values than BHA; however, it points out that the polarizability of 11 (characterized by two carboxylic groups) was influenced by the complexity system and the number of electrons; hence, it is considered the highest polar compound among this series.

Table 2. HOMO and LUMO Energies, Mean Polarizability, and Dipole Moment of Compounds 1–13 at the B3LYP/6-311G(d,p) Level of Theory.
  EHOMO (eV)
ELUMO (eV)
ΔEHOMO–LUMO (eV)
dipole moment (debye)
polarizability (α) (au)
cps MeOH gas MeOH gas MeOH gas MeOH gas MeOH gas
1 –5.89 –5.859 –1.18 –0.933 4.71 4.925 6.535 5.1315 134.839 101.13
2 –5.701 –5.622 –1.647 –1.584 4.053 4.037 7.976 6.3430 289.455 218.926
3 –5.669 –6.143 –1.7 –1.729 3.969 4.414 9.339 7.3883 296.714 225.846
4 –5.892 –5.964 –2.675 –2.764 3.217 3.2 9.692 7.4601 299.938 225.050
5 –6.229 –5.971 –1.404 –1.469 4.825 4.503 3.809 2.7875 203.252 155.961
6 –6.219 –5.934 –1.405 –1.455 4.814 4.478 4.05 3.0199 219.315 168.513
7 –6.191 –5.999 –1.459 –1.488 4.732 4.511 5.227 2.7013 230.721 180.334
8 –6.167 –6.017 –1.367 –1.134 4.8 4.884 10.687 7.7614 244.068 190.842
9 –6.195 –5.952 –1.446 –1.448 4.749 4.503 4.729 3.0260 286.6 219.901
10 –6.181 –5.866 –1.444 –1.445 4.737 4.42 4.954 2.9235 298.64 227.654
11 –6.412 –6.622 –3.46 –3.806 2.952 2.816 3.451 2.0656 426.991 329.818
12 –6.347 –6.329 –2.611 –2.719 3.736 3.61 7.749 5.8543 250.778 190.286
13 –6.411 –6.509 –2.878 –3.097 3.533 3.412 4.273 3.0994 287.514 219.863
BHA –5.627 –5.546 –0.371 –0.228 5.256 5.318 3.156 2.5045 172.023 131.446

2.5.2. Frontier Molecular Orbitals and Reactivity Descriptors

The molecular orbital (MO) analysis of EHOMO and ELUMO is correlated to the chemical reactivity and stability of the compounds, in which molecules can be polarized effectively when small gap energy (ΔEHOMO–LUMO) occurs between HOMO and LUMO. This is attributed to the transfer of electronic charges from donor to withdrawal groups. However, when the ΔEHOMO–LUMO gaps of molecules are very large, the polarization of molecules appears to be less effective and reactive than that of the former (with a smaller gap). As illustrated in Table 2, the results of calculated gaps for 1–13 are in ranges of 2.952–4.825 and 2.816–4.925 eV in methanol and gas phases, respectively, compared to those of BHA (5.20 and 5.32 eV). Figure 5 shows the electron density distribution of HOMO in the gas phase for compounds 14 and 1113, which mainly occurred on (Ar–NH) and (NH–CS–NH–NH−) moieties as the electron transfer zones, while in 510, it distributed mostly on the (NH–CS–NH–NH2) group. LUMO spatial distributions of 14 and 1113 are located on the carboxylic, phenyl hydroxyl, and isoindoline-1,3-dionyl moieties (the electron acceptor zones); however, they are mostly located on −NH–NH– in 510. The −N–NH, −N–NH–CS–NH–, and Ar–OH groups could be the most plausible reaction sites for the free radical attack. Molecules with lower frontier molecular orbital gaps are more polarized.33 Theoretically, the HOMO energy is a valid indicator for scavenging activity rather than LUMO energy.34 All values of global reactivity descriptors are presented in Tables 1 and 2 and reported in the eV unit. The chemical hardness (η) is correlated to the chemical potential with the resistance of change in the number of electrons (molecule stability), whereas S provides an insight into the capacity of an atom to receive an electron (molecule reactivity). In comparison with the calculated S of BHA (0.381 and 0.376 in the methanol and gas phases, respectively), compounds 1, 3, and 510 showed S values ranging between 0.405 and 0.453 in the gas phase, making them more chemically active and easier to donate electrons or H supply. However, the calculated η values (2.207–2.463) in the gas phase reflect their good stabilities in relation to BHA (2.659). The S parameter demonstrates that compounds 1, 3, and 510 are more preferable in the charge-transfer mechanism. For understanding the charge-transfer reaction, the χ and μ parameters were applied to measure the electron escaping tendency. In accordance with the obtained outcomes, compounds 13 and 510 showed the lowest χ values (3.396–3.936) in comparison to 4 and 1113 (4.364–5.214) in the gas phase, indicating their good antioxidant abilities via their ability to donate electrons. Generally, all μ, χ, S, and η descriptor values suggest that compounds 1, 3, and 510 were characterized as more promising targets for electron scavenging reactions, which is in agreement with the experimental assay results.

Figure 5.

Figure 5

Figure 5

HOMO and LUMO molecular orbital plots for 1–13 in the gas phase.

Table 1. Global Reactivity Descriptors of Compounds 113 at the B3LYP/6-311G(d,p) Level of Theory.
  η S ω μ χ
formula = (EHOMO – ELUMO)/2
= 1/η
μ2/2η
= (EHOMO + ELUMO)/2
= −μ
cps MeOH gas MeOH gas MeOH gas MeOH gas MeOH gas
1 2.355 2.463 0.425 0.406 2.653 2.342 –3.535 –3.396 3.535 3.396
2 2.027 2.019 0.493 0.495 3.33 3.215 –3.674 –3.603 3.674 3.603
3 1.984 2.207 0.504 0.453 3.421 3.51 –3.685 –3.936 3.685 3.936
4 1.609 1.6 0.622 0.625 5.704 5.952 –4.284 –4.364 4.284 4.364
5 2.412 2.251 0.415 0.444 3.019 3.073 –3.817 –3.72 3.817 3.72
6 2.407 2.239 0.415 0.447 3.019 3.048 –3.812 –3.694 3.812 3.694
7 2.366 2.256 0.423 0.443 3.091 3.106 –3.825 –3.743 3.825 3.743
8 2.4 2.442 0.417 0.41 2.956 2.618 –3.767 –3.576 3.767 3.576
9 2.375 2.252 0.421 0.444 3.073 3.04 –3.82 –3.7 3.82 3.7
10 2.368 2.21 0.422 0.452 3.068 3.023 –3.812 –3.656 3.812 3.656
11 1.476 1.408 0.678 0.71 8.254 9.655 –4.936 –5.214 4.936 5.214
12 1.868 1.805 0.535 0.554 5.371 5.668 –4.479 –4.524 4.479 4.524
13 1.766 1.706 0.566 0.586 6.105 6.761 –4.644 –4.803 4.644 4.803
BHA 2.628 2.659 0.381 0.376 1.711 1.568 –2.999 –2.887 2.999 2.887

Molecular electrostatic potential (MEP) is a crucial descriptor used to validate the evidence parts of the molecular reactivity system. The MEP three-dimensional surface provides us with information about the position, shape, positive/negative size, and neutral electrostatic potentials.32 Such information highlighted the structural relationship and molecular reactivity toward electrophilic and nucleophilic attacks. In Figure 6, the maximum negative electronic potential (in red color) is the favored site for the electrophilic attack, while the positive electrostatic potential (in blue color) is attracted by radical charged molecules. The high electrostatic potential regions (Figure 6) are observed in the vicinity of the oxygen of the carboxylic group in 113 and in the sulfur of the thiourea group in 510, while a low electropositive potential existed on the H1 of (Ar–NH) of all molecules, H2 of hydrazine (Ar–NH–NH) in molecules of 1 and 510, Ar–OH in 24 and (R–NHCSNH−) in 510. By comparing the blue color code values, +0.104 au in 4, +0.0844 in 2, +0.0844 in 3, +0.08.371 in 8, and +0.08153 in 7 suggest that Ar–NH in 7 and 8 and Ar–OH in 24 are the most favored nucleophilic attacking sites. Moreover, in 510, the codes ranging between +0.08371 and +0.07394 indicate their favored site for nucleophilic interactions.

Figure 6.

Figure 6

Figure 6

Molecular electrostatic potential of 113.

In addition, the H atoms of the hydrazine group are appropriate sites for nucleophiles, whereas the oxygen atom of carboxylic acid is favored for electrophiles.

Free radicals interact with organic molecules by one or more of the following three fundamental mechanisms: HAT, SETPT, and SPLET (Table S2).32 The resulting parameters (BDE, IP, PDE, and PA) shown in Table 3 indicate which radical scavenging mechanism is favored over another and the detailed HAT, SETPT, and SPLET mechanisms are reported.32 To describe the HAT mechanism, BDE is the best consistent thermodynamic parameter. This mechanism includes the transformation of the H atom from the NH– or OH– of the antioxidant compounds to the free radical (Scheme S1). The lowest BDE is the most readily available antioxidant in the highest quantity. As illustrated in Table 3, the results revealed that the N–H and O–H of compounds 3 and 510 have the least BDE values and demonstrated the highest radical scavenging properties, which is in agreement with the experimental results. In the present study, BDEs were calculated in the methanol and gas phases.35 The obtained results revealed that the BDE values in methanol are in a range of 4.83–29.79 kcal/mol, which are higher than those in the gas phase for 510. However, the lowest BDE values in the gas phase are 80.137, 70.67, 67.597, 67.071, 60.863, 67.532, and 66.296 kcal/mol for 1, 3, 57, 9, and 10, respectively (Table 3), with regard to BHA (69.77 kcal/mol). Surprisingly, the N–H bonds of 1, 5, 6, 7, 9, and 10 are still the easiest dissociation bonds in methanol solvent. Based on the calculated BDE values, compounds 3 and 5–10 are considered the most potential antioxidants. The IP and PDE (Table 3) are the most important parameters for the SETPT mechanism.32 In general, lower IPs are more subject to ionization and increase the electron transfer rate between free radicals and antioxidants. By comparison, the IP values of the target compounds were ordered as 3 < 2 < BHA < 4 < 1 < 8 < 7 < 10 < 9 < 6 < 5 < 12 < 13 < 11. The IP values of the investigated compounds 1 and 413 are larger than the PI value of BHA except for 2 and 3. Thus, compounds 2 and 3 are the easiest to transfer electrons and possess one of the highest antioxidant activities. Moreover, compounds 5, 6, 9, 10 and 11 revealed the lowest calculated PDEs of approximately 150–260 kcal/mol. In accordance with these results, compounds 57, 9, and 10 should be the most potental antioxidants.32

Table 3. Calculated BDE, IP, PDE, and PA for 113 at the B3LYP/6-311G(d,p) Level of Theory (in kcal/mol).
    BDE
       
comp. H position MeOH gas PDE IP PA ABTS%
1 Ar–NH 86.65 80.137 266.321 136.189 188.479 74.52 ± 0.11
R–NH 86.338 81.513 274.785 319.852
2 Ar–NH 84.649 79.825 270.859 129.65 306.949 82.28 ± 0.20
Ar–OH 85.763 80.939 271.973 295.383
3 Ar–NH 85.307 72.514 271.775 129.392 307.6 85.19 ± 0.17
Ar–OH 84.136 70.67 270.604 303.835
4 Ar–NH 84.997 72.061 265.59 135.268 303.998 69.23 ± 0.15
Ar–OH 91.062 75.365 271.654 283.123
5 Ar–NH 84.472 67.597 258.693 141.64 305.743 81.54 ± 0.20
R–NH2 89.542 72.672 263.762 307.873
R–NH3 102.028 91.604 276.248 302.757
6 Ar–NH 84.223 67.529 258.77 141.314 305.123 80.60 ± 0.15
R–NH2 89.324 72.693 263.871 307.823
R–NH3 102.32 91.267 276.867 310.686
7 Ar–NH 86.013 67.071 261.054 140.82 304.845 84.34 ± 0.10
R–NH2 92.574 71.943 267.615 305.255
R–NH3 101.349 92.388 276.39 309.637
8 Ar–NH 89.319 74.163 265.171 140.009 309.197 82.49 ± 0.12
R–NH2 90.656 60.863 266.508 299.765
R–NH3 110.757 91.335 286.609 308.775
9 Ar–NH 85.036 67.532 259.785 141.112 304.476 83.41 ± 0.21
R–NH2 91.311 73.181 266.06 305.667
R–NH3 102.635 86.156 277.384 309.372
10 Ar–NH 85.445 66.296 260.46 140.845 304.271 83.22 ± 0.20
R–NH2 91.202 71.583 266.218 305.868
R–NH3 101.797 90.582 276.813 302.004
11 Ar–NH 89.691 77.423 150.281 255.271 301.252 34.77 ± 0.15
12 Ar–NH 89.716 77.505 259.517 146.06 304.715 18.51 ± 0.17
13 Ar–NH 89.824 77.467 258.044 147.64 303.076 46.55 ± 0.16
BHA Ar–OH 82.929 69.77 266.122 132.667 309.53 85.18 ± 0.33

The SPLET mechanism, as shown in Table S2, initiates with the dissociation of the acidic moiety, which can be characterized by the proton affinity (suggested as the possible path to trap the radical in the polar environment); this is followed by an electron transfer to the free radical, at a cost of the electron transfer energy.32 The PA parameter is employed in this regard for elucidation purposes. Thus, lower PA has higher antioxidant capacity. As can be seen from Table 3, the PA values of O–H bonds are lower than those of N–H bonds. For instance, the O–H bonds have the lowest PA values in the methanol phase for 2–4 (295.383, 303.835, and 283.123 kcal/mol, respectively), whereas the N–H bonds have PA values of 305.743, 305.123, 304.845, 304.476, and 304.271 kcal/mol for 5–7, 9, and 10, respectively. Thus, the N–H bonds are also favored for the SPLET mechanism. This variant in activity is mostly associated with NH/OH group numbers that can be transferred to the DPPH free radical. Thus, an increase in hydrogen atoms is accompanied by an increase in scavenging activities. The findings suggested that the studied compounds might demonstrate their antioxidant activity via the three abovementioned antioxidant mechanisms. Due to the polarizability of 1–13, their electron scavenging reactions were obviously influenced by the polar environment, and polarizability could positively enhance the contribution of the SETPT mechanism followed by the SPLET mechanism. An agreement was noticed between some calculated values and the experimental results.

3. Conclusions

Inspired by the antioxidant properties of molecules bearing isothiocyanate and hydrazine core units, targets 113 were evaluated for their antioxidant activities. The results revealed that all investigated compounds showed antioxidant activity, particularly compounds 510 exhibited the highest scavenging effects. The radical scavenging activity of 113 was evaluated theoretically at the B3LYP/6-311++G(d,p) level of theory. The computed global reactivity descriptors suggested that molecules 1 and 510 were more favorable in the electron scavenging reactions. A good correlation was found between the experimental and calculated data. Compounds 3 and 510 had the least BDE values and demonstrated the highest radical scavenging properties. The calculated chemical descriptor values (μ, χ, S, and η) suggested 1, 3, and 510 as more promising targets for electron scavenging reactions. In addition, the HOMO and LUMO energies of targets 113 were calculated using the above DFT set to determine the properties of the target compounds and facilitate their antioxidant mechanisms.

4. Materials and Methods

4.1. Antioxidant Activity Investigation Assays

The investigated compounds 113 were described and characterized in our previous work.27 In accordance with the procedures of Brand-Williams, Oyaizu, Benzie, and Re et al.,3639 the DPPH, reducing power capability, FRAP, and ABTS radical scavenging assays were performed. The detailed methodologies were reported in our previous papers.4042

4.2. DFT Study and Antioxidant Computation Descriptors

By employing the Gaussian 09 program package, all calculations were enforced.20 Compounds structures, radicals, and ions were optimized by DFT, combined with the B3LYP/6-311G(d,p) set.32,43 Assessment of free radical scavenging mechanisms for hydrazinobenzoic acids 113 was accomplished by the computation of various antioxidant descriptors, as presented in Tables S1 and S2.

Acknowledgments

The authors extend their appreciation to Research Supporting Project, King Saud University, Riyadh, Saudi Arabia, for funding this work through grant no. RSP-2021/353.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.1c04772.

  • Initial inputs for geometry optimizations of 113 and BHA (Figure S1); chemical reactivity descriptors (Table S1); antioxidant reaction mechanism and computation of antioxidant descriptors (Table S2); and example of the expected reaction mechanism between the DPPH free radical and compounds 510 (Scheme S1) (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao1c04772_si_001.pdf (562.5KB, pdf)

References

  1. Gülçin İ.; Trofimov B.; Kaya R.; Taslimi P.; Sobenina L.; Schmidt E.; Petrova O.; Malysheva S.; Gusarova N.; Farzaliyev V.; et al. Synthesis of Nitrogen, Phosphorus, Selenium and Sulfur-Containing Heterocyclic Compounds–Determination of Their Carbonic Anhydrase, Acetylcholinesterase, Butyrylcholinesterase and α-Glycosidase Inhibition Properties. Bioorg. Chem. 2020, 103, 104171 10.1016/j.bioorg.2020.104171. [DOI] [PubMed] [Google Scholar]
  2. Kazancı A.; Gök Y.; Kaya R.; Aktaş A.; Taslimi P.; Gülçin İ. Synthesis, Characterization and Bioactivities of Dative Donor Ligand N-Heterocyclic Carbene (NHC) Precursors and Their Ag (I) NHC Coordination Compounds. Polyhedron 2021, 193, 114866 10.1016/j.poly.2020.114866. [DOI] [Google Scholar]
  3. Behçet A.; Aktaş A.; Gök Y.; Kaya R.; Taslimi P.; Gülçin İ. Novel Silver (I) N-heterocyclic Carbene Complexes Bearing 2-(4-hydroxyphenyl) Ethyl Group: Synthesis, Characterization, and Enzyme Inhibition Properties. J. Heterocycl. Chem. 2021, 58, 603–611. 10.1002/jhet.4199. [DOI] [Google Scholar]
  4. Saouli S.; Selatnia I.; Zouchoune B.; Sid A.; Zendaoui S. M.; Bensouici C.; Bendeif E.-E. Synthesis, Spectroscopic Characterization, Crystal Structure, DFT Studies and Biological Activities of New Hydrazone Derivative: 1-(2, 5-Bis ((E)-4-Isopropylbenzylidene) Cyclopentylidene)-2-(2, 4-Dinitrophenyl) Hydrazine. J. Mol. Struct. 2020, 1213, 128203 10.1016/j.molstruc.2020.128203. [DOI] [Google Scholar]
  5. Markova N. V.; Rogojerov M. I.; Angelova V. T.; Vassilev N. G. Experimental and Theoretical Conformational Studies of Hydrazine Derivatives Bearing a Chromene Scaffold. J. Mol. Struct. 2019, 1198, 126880 10.1016/j.molstruc.2019.126880. [DOI] [Google Scholar]
  6. Chaitra T. K.; Mohana K. N.; Gurudatt D. M.; Tandon H. C. Inhibition Activity of New Thiazole Hydrazones towards Mild Steel Corrosion in Acid Media by Thermodynamic, Electrochemical and Quantum Chemical Methods. J. Taiwan Inst. Chem. Eng. 2016, 67, 521–531. 10.1016/j.jtice.2016.08.013. [DOI] [Google Scholar]
  7. Chafai N.; Chafaa S.; Benbouguerra K.; Hellal A.; Mehri M. Synthesis, Spectral Analysis, Anti-Corrosive Activity and Theoretical Study of an Aromatic Hydrazone Derivative. J. Mol. Struct. 2019, 1181, 83–92. 10.1016/j.molstruc.2018.12.073. [DOI] [Google Scholar]
  8. SICAK Y. Investigation of Antioxidant, Anticholinesterase Inhibitory, Tyrosinase Inhibitory and Urease Inhibitory Activities of Some Hydrazone Derivatives. Türk Yaşam Bilim. Derg. 2017, 2, 165–170. [Google Scholar]
  9. Altalhi A. A.; Hashem H. E.; Negm N. A.; Mohamed E. A.; Azmy E. M. Synthesis, Characterization, Computational Study, and Screening of Novel 1-Phenyl-4-(2-Phenylacetyl)-Thiosemicarbazide Derivatives for Their Antioxidant and Antimicrobial Activities. J. Mol. Liq. 2021, 333, 115977 10.1016/j.molliq.2021.115977. [DOI] [Google Scholar]
  10. Badran I.; Tighadouini S.; Radi S.; Zarrouk A.; Warad I. Experimental and First-Principles Study of a New Hydrazine Derivative for DSSC Applications. J. Mol. Struct. 2021, 1229, 129799 10.1016/j.molstruc.2020.129799. [DOI] [Google Scholar]
  11. Al-Salahi R.; Marzouk M.; Ashour A. E.; Alswaidan I. Synthesis and Antitumor Activity of 1,2,4-Triazolo[1,5-a]Quinazolines. Asian J. Chem. 2014, 26, 2173–2176. 10.14233/ajchem.2014.16849. [DOI] [Google Scholar]
  12. Polović S.; LJoljić Bilić V.; Budimir A.; Kontrec D.; Galić N.; Kosalec I. Antimicrobial Assesment of Aroylhydrazone Derivatives in Vitro. Acta Pharm. 2019, 69, 277–285. 10.2478/acph-2019-0020. [DOI] [PubMed] [Google Scholar]
  13. Debnath U.; Mukherjee S.; Joardar N.; Babu S. P. S.; Jana K.; Misra A. K. Aryl Quinolinyl Hydrazone Derivatives as Anti-Inflammatory Agents That Inhibit TLR4 Activation in the Macrophages. Eur. J. Pharm. Sci. 2019, 134, 102–115. 10.1016/j.ejps.2019.04.016. [DOI] [PubMed] [Google Scholar]
  14. Narayanan S.; Gupta P.; Nazim U.; Ali M.; Karadkhelkar N.; Ahmad M.; Chen Z.-S. Anti-Cancer Effect of Indanone-Based Thiazolyl Hydrazone Derivative on Colon Cancer Cell Lines. Int. J. Biochem. Cell Biol. 2019, 110, 21–28. 10.1016/j.biocel.2019.02.004. [DOI] [PubMed] [Google Scholar]
  15. Cedrowski J.; Dąbrowa K.; Przybylski P.; Krogul-Sobczak A.; Litwinienko G. Antioxidant Activity of Two Edible Isothiocyanates: Sulforaphane and Erucin Is Due to Their Thermal Decomposition to Sulfenic Acids and Methylsulfinyl Radicals. Food Chem. 2021, 353, 129213 10.1016/j.foodchem.2021.129213. [DOI] [PubMed] [Google Scholar]
  16. Khodja I. A.; Boulebd H.; Bensouici C.; Belfaitah A. Design, Synthesis, Biological Evaluation, Molecular Docking, DFT Calculations and in Silico ADME Analysis of (Benz) Imidazole-Hydrazone Derivatives as Promising Antioxidant, Antifungal, and Anti-Acetylcholinesterase Agents. J. Mol. Struct. 2020, 1218, 128527 10.1016/j.molstruc.2020.128527. [DOI] [Google Scholar]
  17. Gulcin İ. Antioxidants and Antioxidant Methods: An Updated Overview. Arch. Toxicol. 2020, 94, 651–715. 10.1007/s00204-020-02689-3. [DOI] [PubMed] [Google Scholar]
  18. Taslimi P.; Gulçin İ. Antioxidant and Anticholinergic Properties of Olivetol. J. Food Biochem. 2018, 42, e12516 10.1111/jfbc.12516. [DOI] [Google Scholar]
  19. Zhong Z.; Zhong Z.; Xing R.; Li P.; Mo G. The Preparation and Antioxidant Activity of 2-[Phenylhydrazine (or Hydrazine)-Thiosemicarbazone]-Chitosan. Int. J. Biol. Macromol. 2010, 47, 93–97. 10.1016/j.ijbiomac.2010.05.016. [DOI] [PubMed] [Google Scholar]
  20. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Farkas O.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. et al. Gaussian 09, Revision A, 2009; 1.
  21. Dehghani Z.; Khoshneviszadeh M.; Khoshneviszadeh M.; Ranjbar S. Veratric Acid Derivatives Containing Benzylidene-Hydrazine Moieties as Promising Tyrosinase Inhibitors and Free Radical Scavengers. Bioorg. Med. Chem. 2019, 27, 2644–2651. 10.1016/j.bmc.2019.04.016. [DOI] [PubMed] [Google Scholar]
  22. Al-Salahi R.; Anouar E.-H.; Marzouk M.; Taie H. A. A.; Abuelizz H. A. Screening and Evaluation of Antioxidant Activity of Some 1, 2, 4-Triazolo[1, 5-a]Quinazoline Derivatives. Future Med. Chem. 2018, 10, 379–390. 10.4155/fmc-2017-0224. [DOI] [PubMed] [Google Scholar]
  23. Al-Salahi R.; Taie H. A. A.; Bakheit A. H.; Marzouk M.; Almehizia A. A.; Herqash R.; Abuelizz H. A. Antioxidant Activities and Molecular Docking of 2-Thioxobenzo[g]Quinazoline Derivatives. Pharmacol. Rep. 2019, 71, 695–700. 10.1016/j.pharep.2019.04.003. [DOI] [PubMed] [Google Scholar]
  24. Almehizia A. A.; Abuelizz H. A.; Taie H. A. A.; ElHassane A.; Marzouk M.; Al-Salahi R. Investigation the Antioxidant Activity of Benzo[g]Triazoloquinazolines Correlated with a DFT Study. Saudi Pharm. J. 2019, 27, 133–137. 10.1016/j.jsps.2018.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Khan I.; Ibrar A.; Abbas N.; Saeed A. Recent Advances in the Structural Library of Functionalized Quinazoline and Quinazolinone Scaffolds: Synthetic Approaches and Multifarious Applications. Eur. J. Med. Chem. 2014, 76, 193–244. 10.1016/j.ejmech.2014.02.005. [DOI] [PubMed] [Google Scholar]
  26. Tetere Z.; Zica̅ne D.; Ra̅viņa I.; Mieriņa I.; Rijkure I. Synthesis and Antioxidant Properties of Novel Quinazoline Derivatives. Mater. Sci. Appl. Chem. 2014, 30, 51–54. 10.7250/msac.2014.009. [DOI] [Google Scholar]
  27. Abuelizz H. A.; Awad H. M.; Marzouk M.; Nasr F. A.; Bakheit A. H.; Naglah A. M.; Al-Shakliah N. S.; Al-Salahi R. Exploiting the 4-Hydrazinobenzoic Acid Moiety for the Development of Anticancer Agents: Synthesis and Biological Profile. Bioorg. Chem. 2020, 102, 104098 10.1016/j.bioorg.2020.104098. [DOI] [PubMed] [Google Scholar]
  28. Cakmak K. C.; Gülçin İ. Anticholinergic and Antioxidant Activities of Usnic Acid-An Activity-Structure Insight. Toxicol. Rep. 2019, 6, 1273–1280. 10.1016/j.toxrep.2019.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Eruygur N.; Koçyiğit U. M.; Taslimi P.; Ataş M.; Tekin M.; Gülçin İ. Screening the in Vitro Antioxidant, Antimicrobial, Anticholinesterase, Antidiabetic Activities of Endemic Achillea Cucullata (Asteraceae) Ethanol Extract. S. Afr. J. Bot. 2019, 120, 141–145. 10.1016/j.sajb.2018.04.001. [DOI] [Google Scholar]
  30. Öztaşkın N.; Kaya R.; Maraş A.; Şahin E.; Gülcin İ.; Göksu S. Synthesis and Characterization of Novel Bromophenols: Determination of Their Anticholinergic, Antidiabetic and Antioxidant Activities. Bioorg. Chem. 2019, 87, 91–102. 10.1016/j.bioorg.2019.03.010. [DOI] [PubMed] [Google Scholar]
  31. Savegnago L.; do Sacramento M.; Brod L. M. P.; Fronza M. G.; Seus N.; Lenardao E. J.; Paixão M. W.; Alves D. Phenylselanyl-1 H-1, 2, 3-Triazole-4-Carbonitriles: Synthesis, Antioxidant Properties and Use as Precursors to Highly Functionalized Tetrazoles. RSC Adv. 2016, 6, 8021–8031. 10.1039/C5RA22445D. [DOI] [Google Scholar]
  32. Farrokhnia M. Density Functional Theory Studies on the Antioxidant Mechanism and Electronic Properties of Some Bioactive Marine Meroterpenoids: Sargahydroquionic Acid and Sargachromanol. ACS Omega 2020, 5, 20382–20390. 10.1021/acsomega.0c02354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. El Bahnasawy R. M.; El Shereafy E.; Kashar T. I. Thermal and Temperature Dependence of Electrical Conductivity Studies on Zn, Cd and Hg Hydrazone Complexes. J. Therm. Anal. 1993, 39, 65–74. 10.1007/BF02235447. [DOI] [Google Scholar]
  34. Al-Majedy Y. K.; Al-Amiery A. A.; Kadhum A. A. H.; Mohamad A. B. Antioxidant Activities of 4-Methylumbelliferone Derivatives. PLoS One 2016, 11, e0156625 10.1371/journal.pone.0156625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Vo Q. V.; Nam P. C.; Van Bay M.; Thong N. M.; Cuong N. D.; Mechler A. Density Functional Theory Study of the Role of Benzylic Hydrogen Atoms in the Antioxidant Properties of Lignans. Sci. Rep. 2018, 8, 12361 10.1038/s41598-018-30860-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Brand-Williams W.; Cuvelier M.-E.; Berset C. Use of a Free Radical Method to Evaluate Antioxidant Activity. LWT-Food Sci. Technol. 1995, 28, 25–30. 10.1016/S0023-6438(95)80008-5. [DOI] [Google Scholar]
  37. Oyaizu M. Antioxidative Activities of Browning Reaction Prepared from Glucosamine. Jpn. J. Nutr. Diet. 1986, 44, 307–315. 10.5264/eiyogakuzashi.44.307. [DOI] [Google Scholar]
  38. Benzie I. F. F.; Strain J. J. The Ferric Reducing Ability of Plasma (FRAP) as a Measure of “Antioxidant Power”: The FRAP Assay. Anal. Biochem. 1996, 239, 70–76. 10.1006/abio.1996.0292. [DOI] [PubMed] [Google Scholar]
  39. Re R.; Pellegrini N.; Proteggente A.; Pannala A.; Yang M.; Rice-Evans C. Antioxidant Activity Applying an Improved ABTS Radical Cation Decolorization Assay. Free Radical Biol. Med. 1999, 26, 1231–1237. 10.1016/S0891-5849(98)00315-3. [DOI] [PubMed] [Google Scholar]
  40. Abuelizz H. A.; Taie H. A. A.; Marzouk M.; Al-Salahi R. Synthesis and Antioxidant Activity of 2-Methylthio-Pyrido [3, 2-e][1, 2, 4] Triazolo[1, 5-a]Pyrimidines. Open Chem. 2019, 17, 823–830. 10.1515/chem-2019-0092. [DOI] [Google Scholar]
  41. Al-Salahi R.; Taie H. A. A.; Bakheit A. H.; Marzouk M.; Almehizia A. A.; Herqash R.; Abuelizz H. A. Antioxidant Activities and Molecular Docking of 2-Thioxobenzo[g]Quinazoline Derivatives. Pharmacol. Rep. 2019, 71, 695–700. 10.1016/j.pharep.2019.04.003. [DOI] [PubMed] [Google Scholar]
  42. Abuelizz H. A.; Anouar E.; Marzouk M.; Taie H. A. A.; Ahudhaif A.; Al-Salahi R. DFT Study and Radical Scavenging Activity of 2-Phenoxypyridotriazolo Pyrimidines by DPPH, ABTS, FRAP and Reducing Power Capacity. Chem. Pap. 2020, 74, 2893–2899. 10.1007/s11696-020-01126-0. [DOI] [Google Scholar]
  43. Galano A.; Raúl Alvarez-Idaboy J. Computational Strategies for Predicting Free Radical Scavengers’ Protection against Oxidative Stress: Where Are We and What Might Follow?. Int. J. Quantum Chem. 2019, 119, e25665 10.1002/qua.25665. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao1c04772_si_001.pdf (562.5KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES