Next Article in Journal
Environmentally Sustainable Achiral and Chiral Chromatographic Analysis of Amino Acids in Food Supplements
Next Article in Special Issue
Continuous-Flow Chemistry and Photochemistry for Manufacturing of Active Pharmaceutical Ingredients
Previous Article in Journal
Cytotoxic and Pro-Apoptotic Effects of Leaves Extract of Antiaris africana Engler (Moraceae)
Previous Article in Special Issue
Novel Complexes of 3-[3-(1H-Imidazol-1-yl)propyl]-3,7-diaza-bispidines and β-Cyclodextrin as Coatings to Protect and Stimulate Sprouting Wheat Seeds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of 4-Aminopyrazol-5-ols as Edaravone Analogs and Their Antioxidant Activity

by
Yanina V. Burgart
1,
Galina F. Makhaeva
2,
Olga P. Krasnykh
3,
Sophia S. Borisevich
4,
Natalia A. Agafonova
1,
Nadezhda V. Kovaleva
2,
Natalia P. Boltneva
2,
Elena V. Rudakova
2,
Evgeny V. Shchegolkov
1,
Galina A. Triandafilova
3,
Denis A. Gazizov
1,
Olga G. Serebryakova
2,
Maria V. Ulitko
5,
Sergey L. Khursan
4,
Victor I. Saloutin
1 and
Rudy J. Richardson
6,7,8,9,*
1
Postovsky Institute of Organic Synthesis of the Ural Branch of the Russian Academy of Science, S. Kovalevskoi St., 22, Ekaterinburg 620108, Russia
2
Institute of Physiologically Active Compounds at Federal Research Center of Problems of Chemical Physics and Medicinal Chemistry, Russian Academy of Sciences (IPAC RAS), Severny proezd 1, Chernogolovka 142432, Russia
3
Scientific and Educational Center for Applied Chemical-Biological Research, Perm National Research Polytechnic University, Komsomolsky Av., 29, Perm 614990, Russia
4
Ufa Institute of Chemistry of Russian Academy of Science, Octyabrya Av., 71, Ufa 450078, Russia
5
Institute of Natural Sciences and Mathematics of the Ural Federal University Named after the First President of Russia B. N. Yeltsin, Lenina Av., 51, Ekaterinburg 620083, Russia
6
Department of Environmental Health Sciences, University of Michigan, Ann Arbor, MI 48109, USA
7
Department of Neurology, University of Michigan, Ann Arbor, MI 48109, USA
8
Center of Computational Medicine and Bioinformatics, University of Michigan, Ann Arbor, MI 48109, USA
9
Michigan Institute for Computational Discovery and Engineering, University of Michigan, Ann Arbor, MI 48109, USA
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(22), 7722; https://doi.org/10.3390/molecules27227722
Submission received: 21 October 2022 / Revised: 2 November 2022 / Accepted: 4 November 2022 / Published: 9 November 2022
(This article belongs to the Special Issue Synthesis of Bioactive Compounds)

Abstract

:
One of the powerful antioxidants used clinically is Edaravone (EDA). We synthesized a series of new EDA analogs, 4-aminopyrazol-5-ol hydrochlorides, including polyfluoroalkyl derivatives, via the reduction of 4-hydroxyiminopyrazol-5-ones. The primary antioxidant activity of the compounds in comparison with EDA was investigated in vitro using ABTS, FRAP, and ORAC tests. In all tests, 4-Amino-3-pyrazol-5-ols were effective. The lead compound, 4-amino-3-methyl-1-phenylpyrazol-5-ol hydrochloride (APH), showed the following activities: ABTS, 0.93 TEAC; FRAP, 0.98 TE; and ORAC, 4.39 TE. APH and its NH-analog were not cytotoxic against cultured normal human fibroblasts even at 100 μM, in contrast to EDA. According to QM calculations, 4-aminopyrazolols were characterized by lower gaps, IP, and η compared to 4-hydroxyiminopyrazol-5-ones, consistent with their higher antioxidant activities in ABTS and FRAP tests, realized by the SET mechanism. The radical-scavenging action evaluated in the ORAC test occurred by the HAT mechanism through OH bond breaking in all compounds, directly dependent on the dissociation energy of the OH bond. All the studied compounds demonstrated the absence of anticholinesterase activity and moderate inhibition of CES by some 4-aminopyrazolols. Thus, the lead compound APH was found to be a good antioxidant with the potential to be developed as a novel therapeutic drug candidate in the treatment of diseases associated with oxidative stress.

Graphical Abstract

1. Introduction

Oxidative stress is a crucial factor in the development of serious pathological conditions including cancer, aging, atherosclerosis, rheumatoid arthritis, cardiovascular and autoimmune diseases, and neurodegenerative disorders. It is characterized by an imbalance between the formation and neutralization of free radicals arising from the dysfunction of endogenous antioxidant protection systems. This redox abnormality leads to the critical accumulation of reactive oxygen and nitrogen species (ROS and RNS, respectively), such as peroxides and free radicals that damage cellular proteins, lipids, and nucleic acids. Antioxidants and radical scavengers play vital defensive roles by modulating the concentrations of ROS and RNS. Consequently, they can reduce the occurrence of disorders associated with oxidative stress [1,2].
The pyrazole core is considered to be a unique pharmacophore for the design of promising antioxidants [3,4,5]. One of the powerful antioxidants and neuroprotective drugs in clinical practice is 3-methyl-1-phenyl-2-pyrazolin-5-one (Edaravone, EDA, Radicut, Radicava). This prominent drug was originally used to treat brain infarction (stroke) [6,7,8] and was subsequently approved for the treatment of amyotrophic lateral sclerosis [9,10,11]. In addition, EDA can be used to treat Alzheimer’s disease, glaucoma, pulmonary diseases, cardiovascular dysfunctions, and medical chronic kidney damage, and to reduce specific side effects in the treatment of cancer [12,13,14]. Broadly speaking, EDA can be used to treat a range of diseases involving oxidative stress, and it has no severe adverse effects [12,15,16].
EDA has been found to be active in quenching free radicals, because the phenolic hydroxyl (exhibiting a propensity for scavenging free radicals) is generated from the tautomerization of its C=O group [17,18,19,20,21]. EDA effectively binds hydroxyl radicals by a consistent mechanism for electron–proton transfer [22,23] enabling it to react with alkoxyl radicals, superoxide anions, peroxyl radicals of lipids, peroxynitrite, and nitrogen(II) oxide [18,20,24]. EDA’s scavenging activity against multiple free radical species is as robust as other known potent antioxidants such as uric acid, glutathione, and Trolox [19].
Among analogs of EDA, 3-methyl-1-(pyridin-2-yl)-5-pyrazolone had the highest ability to scavenge radicals due to the increase in its active anion form stabilized by an intramolecular hydrogen bond in [25]. EDA derivatives bearing the NO-donor group in an aryl substituent showed different degrees of balance between antioxidant and vasodilating activity in vitro in [26,27]. Analogs having thiazolyl residues also demonstrated pronounced antioxidant properties in the ABTS test in [28]. Among 4-aryl- and 4-carboxyamide-MeO-derivatives of EDA, compounds with antioxidant properties in vitro were found in [29]. Furthermore, in another study, a series of DL-3-n-butylphthalide-EDA hybrids were synthesized as novel dual inhibitors of β-amyloid aggregation and monoamine oxidases for potential application as Alzheimer’s disease therapeutics, and these conjugates also possessed high antioxidant activity [30].
Thus, the EDA pyrazolone structure can be considered as a promising scaffold for the design of new effective antioxidants. This encouraged us to create new 4-amino-substituted EDA analogs and study their antioxidant properties. We believe that pyrazoles combining amino and hydroxyl groups in their structures should be promising as antioxidants.
For the synthesis of the 4-amino-5-alkoxy-3-methyl-1-phenylpyrazole hydrochloride, the use of reduction of the 4-nitro group in 5-RO-pyrazoles under the action of SnCl2 in concentrated HCl was described in [31]. Furthermore, 1-(sulfonated phenyl)-4-amino-5-pyrazolones were obtained through the reduction of hydroxyimino derivatives by zinc in a mixture of hydrochloric and acetic acids in [32] and by catalytic hydrogenation under the action of Pd/C in [33]. For the synthesis of N-glycoconjugated 4-amino-3-methyl-5-pyrazolones, the reaction was carried out with zinc in methanol in the presence of NH4Cl in [34]. In addition, a method was proposed for the synthesis of 3-substituted 4-aminopyrazol-5-ol hydrochlorides via cyclization of 2-amino-3-oxo esters with hydrazine in [35]. However, it should be noted that although the 4-amino derivative of EDA and its NH-unsubstituted analog were obtained in this way, there was no information on their structures.
Data on polyfluoroalkyl-containing 4-aminopyrazolols are limited to two examples describing the synthesis of 4-alkylamino-3-trifluoromethylpyrazol-5-ols via the recyclization of the relatively inaccessible compound 4-trifluoroacetyl-l,3-oxazolium-5-olate [36] or 5-trifluoromethyl-4-ethoxycarbonyl-1,3-oxazole [35] under the action of hydrazines. However, fluorinated derivatives, including pyrazoles, are undoubtedly promising as pharmacological agents due to the presence of fluorine atoms that change the physical, chemical, and biological properties of organic molecules [37,38,39,40,41,42,43,44]. For example, incorporation of fluorine atoms into drugs can increase their metabolic stability and lipophilicity, thereby facilitating penetration of the molecules through biological membranes [45].
Herein, we focused on the synthesis of 4-amino-substituted EDA derivatives, including polyfluoroalkyl-containing analogs, and the study of their primary antioxidant activity in comparison with EDA, CF3-EDA, and Trolox using ABTS, FRAP, and ORAC tests. To explain the factors influencing antioxidant activity, we performed quantum-chemical calculations. Taking into account the literature data on the ability of pyrazole derivatives to inhibit acetylcholinesterase [46,47,48,49,50] and human carboxylesterase-1 [51], we also investigated the esterase profiles of the new compounds, i.e., their inhibitory activity against acetylcholinesterase (AChE), butyrylcholinesterase (BChE), and carboxylesterase (CES). In addition, the effect of the compounds on cell viability was evaluated in normal human dermal fibroblast cultures.

2. Results and Discussion

2.1. Chemistry

For the synthesis of polyfluoroalkyl-containing 4-aminopyrazol-5-ols, we used the reduction of 4-hydroxyiminopyrazol-5-ones 1aj, for which we have recently proposed convenient synthesis methods [52].
First, we tried to carry out reduction of 4-hydroxyiminopyrazolones 1a,b with zinc in acetic acid, according to our successful method for the synthesis of 4-amino-3-trifluoromethyl-5-alkyl[(het)aryl]pyrazoles from 4-nitrosopyrazoles [53]. However, this method did not allow us to obtain the expected 4-aminopyrazol-5-ols, because bis-[5-hydroxy-1-phenyl-3-(polyfluoroalkyl)-1H-pyrazol-4-yl]imines 2a,b were isolated from these reactions as a result of crosslinking of two pyrazolone molecules by the amino group (Scheme 1). Similar transformations have been described for non-fluorinated analogs [35,54]. Although the mechanism of formation of these compounds is unclear, it can be assumed that the free amines are unstable and convert to bis-pyrazoles 2a,b due to oxidation under the action of oxygen in the air [55].
The bis-pyrazole structure of compound 2a was confirmed by XRD (Figure 1). In its crystal structure, the symmetry axis passing through the exocyclic nitrogen atom N3 and the proton H1 of the hydroxyl group can be marked. The two pyrazole molecules are mirror images of each other and generate a “crab”-like form. The eight-membered pseudocycle O1C5C4N3C4C5O1H1 is practically flat, with the O1 oxygen atom deviating from the plane by 0.353 Å. The compound is characterized by delocalization of double bonds (the lengths of C4N3 and N4C3 are equal to 1.312 Å), and the proton H1 of the hydroxyl group is delocalized between the two O1 oxygen atoms, with the distance O1H1 equal to H1O1 (1.208 Å).
When adding acetic anhydride to the reducing mixture Zn/AcOH, reduction of the hydroxyimine group of pyrazolone 1a was accompanied by acylation of the resulting amino function to yield 4-acylaminopyrazol-5-ol 3a (Scheme 1).
4-Aminopyrazol-5-ols 4ac,fi, as hydrochlorides, were synthesized by reducing 4-hydroxyiminopyrazolones 1ac,fi under the action of tin(II) chloride in concentrated HCl (Scheme 1). However, the use of this synthetic protocol was not effective for the preparation of NH-unsubstituted 4-amino-3-CF3-pyrazol-5-ol 4e, and catalytic hydrogenation in EtOH-HCl in the presence of Pd/C was used for its synthesis. The application of such a system facilitates the isolation of amine salts, as shown in the preparation of compound 4c.
The same method was used to prepare the 4-amino derivative of EDA 4d and its NH-unsubstituted analog 4j from the corresponding 4-hydroxyiminopyrazol-5-ones 1d,j by catalytic reduction (Scheme 1). Compounds 4d,j were characterized by IR along with 1H and 13C NMR spectroscopy, because spectral data for these compounds were lacking [35].
Attempts to carry out catalytic hydrogenation of compounds 1a,b,d under acid-free conditions led to bis-pyrazoles 2a,b or to the resinification of the reaction mass. It can be concluded that 4-aminopyrazol-5-ol as a free base is unstable.
The structure of 4-aminopyrazoles 4aj was confirmed by IR and NMR spectroscopy. The NMR spectra of these compounds recorded in DMSO-d6 are characterized by revealing one hydroxyl form that was confirmed by the presence in the 13C NMR spectra of compounds 4 of carbon atom signals of =C-OH at δ 147–154 ppm. According to NMR spectroscopy data, aminopyrazole hydrochlorides 4a,b (form AH) standing in DMSO solutions were able to convert to ammonium salts (form AS) (Scheme 1). Thus, in the 1H NMR spectrum of hydrochlorides 4a,b, the proton signals of the amino group were broad low-field signals or absent due to deuterium exchange with solvent water, while the spectra after standing of a sample for 6–12 h contain triplet signals of protons of NH3+ groups of ammonium salts. In addition, the content of the AS form increases with elongation of the polyfluoroalkyl substituent. The AH:AS ratio was 77:23 (for 4a) and 40:60 (for 4b). The 13C NMR spectra of compounds 4a,b also contained two sets of signals characterized by the multiplet broadening of the signals of carbon atoms for the ammonium salt and by singlet signals of the same carbons for the hydrochloride. However, compound 4c bearing a nonafluorobutyl residue existed in DMSO solution only as AS, while the methyl-containing analog 4d occurred as AH. All N-unsubstituted aminopyrazoles 4ej were revealed to adopt only the AH form.

2.2. Evaluation of Antioxidant Potential of Pyrazoles 1 and 4

For studying the primary antioxidant activity of compounds 1 and 4, three different methods were selected: (1) the FRAP test, which exclusively evaluates the single electron transfer (SET) mechanism; (2) the ORAC-FL test, evaluating exclusively the hydrogen atom transfer (HAT) mechanism; and (3) the ABTS test, which reflects either HAT or SET mechanisms as well as their combination. For all methods, Trolox was used as the reference antioxidant: antioxidant activity of the compounds was presented relative to the activity of Trolox. Quercetin was used as a positive control. The data are presented in Table 1.
The ABTS assay is based on the direct quenching of the ABTS cation radical (ABTS•+) by antioxidants. The assay is carried out by spectrophotometric determination of a decrease in absorbance of a stable dark green ABTS•+ solution after its interaction with an antioxidant compound [56]. The measurements were performed as previously described in detail [57]. The results are expressed as TEAC values (Trolox equivalent antioxidant capacity) and IC50 values (compound concentration required for 50% reduction of the ABTS radical).
The FRAP (ferric reducing antioxidant power) assay measures the ability of antioxidants to reduce the ferric 2,4,6-tripyridyl-s-triazine complex [Fe(TPTZ)2]3+ to the intensely blue ferrous complex [Fe(TPTZ)2]2+ in an acidic medium [58]. The measurements were performed as previously described in detail [59]. The results are expressed as TE (Trolox equivalents).
In the oxygen radical absorbance capacity (ORAC-FL) assay, 2,2′-azobis-(2-methylpropionamidine) dihydrochloride (AAPH) is used as a free peroxyl radical generator, and fluorescein (FL) is used as a fluorescent probe [60]. The method is based on measuring the decrease in the intensity of fluorescence with time, which characterizes the degree of decay of the fluorescent probe under the influence of peroxyl radicals. In the presence of antioxidants, the degree of decay of the fluorescent probe decreases and, accordingly, the fluorescence intensity increases. The results are expressed as TE (Trolox equivalents).

2.2.1. ABTS Assay

As shown in Table 1, fluorine-containing NH-unsubstituted 4-hydroxyiminopyrazol-5-ones 1ei demonstrated moderate radical-binding activity in the ABTS test, which was practically independent of the RF substituent structure (TEAC = 0.13–0.23). The introduction of a phenyl substituent into position 1 of the pyrazole ring resulted in a sharp decrease in the antiradical activity of compounds 1ac regardless of the RF structure: 1e vs. 1a, (RF = CF3); 1g vs. 1b, 1i vs. 1c (RF = C4F9).
The replacement of the fluorine-containing radical with the methyl group at position 3 promoted an increase in the radical-binding activity of the phenyl-containing derivative (1d vs. 1a) and did not affect the activity of the NH-unsubstituted derivative (1j vs. 1e). The latter compounds exhibited an equally moderate ABTS•+-binding activity.
In contrast to 4-hydroxyiminopyrazol-5-ones 1, aminopyrazoles 4 showed high radical-binding activity in the ABTS test at the level of Trolox and EDA. The ABTS-binding activity of 4-unsubstituted 3-trifluoromethyl-1-phenylpyrazol-5-ol (CF3-EDA) was found to be equal to that of EDA.
In addition, the introduction of a Ph substituent to position 1 of the pyrazole ring in 4-amino-3-RF-pyrazolols 4 retained their high antioxidant activity.
The elongation of the polyfluoroalkyl radical in NH-unsubstituted 4-aminopyrazoles 4 resulted in a gradual decrease in the ABTS-binding activity in the following order: 4e > 4g~4h > 4i. The activity was halved in the transition from CF3-pyrazole 4e to the C4F9 analog 4i.
Similarly to 4-hydroxyiminopyrazolones 1e,j, the replacement of the CF3 group by a Me residue at position 3 fostered an increase in the ABTS•+-binding activity of aminopyrazolol 4d in the presence of a Ph substituent at position 1 (4d vs. 4a) and retention of high activity in the case of NH-unsubstituted derivatives (4j vs. 4e).
Thus, aminopyrazoles 4c,d,e,j exhibited the greatest activity in the ABTS test and had TEAC values at the level of Trolox, EDA, and CF3-EDA. It is also noteworthy that our ABTS test data on the antioxidant activity of EDA matched the values reported in the literature [61,62].

2.2.2. FRAP Assay

Compounds 1 and 4 demonstrated structure–activity relationships in the FRAP assay close to the results from the ABTS test, although in some cases they were less pronounced.
Thus, 4-hydroxyiminopyrazolones 1 were either inactive or exhibited very low activity compared to the standard antioxidant Trolox. The introduction of a phenyl substituent into position N1 reduced ferric reducing activity and, correspondingly, antioxidant activity: 1a vs. 1e (R = CF3), 1b vs. 1g (R = C2F5), 1d vs. 1j (R = Me).
As in the ABTS test, 4-aminopyrazolols 4 were significantly more active as antioxidants compared to 4-hydroxyiminopyrazolones 1. As shown in Table 1, most of the compounds of this group showed high activity at the level of the standard antioxidant Trolox.
The introduction of a Ph-substituent at position 1 of the pyrazole ring of compounds 4 retained their high antioxidant activity compared to unsubstituted ones. Activity increased for compounds with RF = C4F9 (4c was more active than 4i).
Similar to the results of the ABTS test, the replacement of the CF3 group with CH3 at position 3 promoted an increase in ferric reducing activity in the presence of a Ph substituent at position 1 (4a vs. 4d) and maintained high activity in the case of the NH-unsubstituted derivatives (4e vs. 4j).
Interestingly, according to literature data [25], CF3-EDA was 3 times less active compared to EDA in the OH radical binding test (hydroxyl radical scavenging activity). Meanwhile, in the ABTS test, the activity of CF3-EDA was at the level of EDA and Trolox, and in the FRAP test, its ferric reducing activity was only 20% lower than that of EDA (Table 1).
Thus, Ph-substituted (4a,c,d) and NH-unsubstituted (4e,g,h,j) 4-aminopyrazoles showed high activity in the FRAP test. In addition, compounds 4c,d,e,j showed significant activity in the ABTS test. Moreover, the ferric reducing activity of compounds 4a,c,d,e,g,h,j was at the level of Trolox and exceeded the activity of EDA or CF3-EDA.

2.2.3. ORAC Assay

The ORAC assay utilizes a biologically relevant radical source, 2,2′-azobis(2-amidinopropane) dihydrochloride (AAPH)-derived peroxyl radicals. Thus, it is a more physiologically relevant assay than ABTS and FRAP.
As shown in Table 1, overall, the results of the ABTS and FRAP tests agree with the ORAC-FL data; i.e., all compounds that were active in the first two tests showed a pronounced antioxidant effect in the ORAC test. However, there were also differences. In particular, the methyl analog 1d showed activity at the Trolox level (TE = 1.07) in the ORAC test. In addition, NH-unsubstituted 4-hydroxyiminopyrazol-5-ones 1e,h,i,j, which were inactive or low-active in the FRAP test or exhibited moderate activity in the ABTS test, showed a pronounced radical binding effect in the ORAC test. This effect was higher for compounds with a longer polyfluoroalkyl substituent (1e < 1h = 1i).
Structure–activity analysis for results in the ORAC test showed that all compounds in the series of 4-amino derivatives 4 exhibited high antioxidant activity, in contrast to 4-hydroxyimino derivatives 1. The introduction of an N-phenyl substituent into position 1 significantly enhanced the effect (4a vs. 4e, 4b vs. 4g, 4c vs. 4i, 4d vs. 4j).
In each of the subgroups 1ad, 1ej, and 4ad, 4ej, compounds 1d,j and 4d containing a methyl group instead of a polyfluoroalkyl group exhibited higher activity. The most active among the aminopyrazolols 4 were compounds 4a (R1 = CF3, R2 = Ph, 3.49 TE) and 4d (R1 = Me, R2 = Ph, 4.39 TE).
In particular, the lead compound 4d was more active than EDA (3.71 TE) and CF3-EDA (3.89 TE).
IC50 values were determined for compounds 4a,d,e, EDA, CF3-EDA, and Trolox (Table 1). The obtained IC50 values were 3–5 times lower than the IC50 of Trolox, again confirming the high antioxidant activity of the tested pyrazole derivatives.
Thus, the results of the ABTS and FRAP tests show that 4-hydroxyiminopyrazolones 1 had moderate or very weak activity (HN-derivatives 1ej were more active than Ph-N-analogs 1ad), while 4-aminopyrazolols 4 demonstrated a pronounced antioxidant effect comparable to Trolox (additionally, HN- and PhN-derivatives 4ad and 4ej had close activity). The most active compounds in both of these tests were the 4-aminopyrazolols 4d,e,j. In contrast, HN-unsubstituted 4-hydroxyiminopyrazolones 1ej and all 4-aminopyrazolols 4aj were active in the ORAC test. The increase in activity of 4-aminopyrazolol 4 was promoted by the introduction of a phenyl substituent and the replacement of a fluoroalkyl substituent with a methyl group. The lead compound 4d was more active in all tests than EDA and CF3-EDA.

2.3. Assessment of Esterase Profile of Pyrazoles 1aj and 4aj

Given the available literature data on the ability of pyrazole derivatives to inhibit acetylcholinesterase (AChE) [46,47,48,49,50] and human carboxylesterase-1 (CES) [51], we performed an investigation of the esterase profile for new compounds 1aj and 4aj, including an evaluation of their inhibitory effect on the standard enzymes set: human erythrocyte AChE, equine serum butyrylcholinesterase (BChE), and porcine liver CES (Table 2). The applicability of this set of enzymes was shown by us earlier [63,64,65,66].
It was found that 4-hydroxyminopyrazolones 1aj did not substantially inhibit any of the esterases, while aminopyrazolols 4a,c,d,h,i showed moderate anticarboxylesterase activity with IC50 values in the range 10–98 μM, with a maximum activity of 10 and 22 μM for compounds 4c and 4i, respectively, which had a nonafluorobutyl substituent. These data indicate the absence of anticholinesterase side effects, with the potential therapeutic application of these compounds and some probability of undesirable drug interactions from the use of aminopyrazolols 4 at high doses, owing to the inhibition of CES, which participates in the primary metabolism of numerous drugs with ester groups.

2.4. Quantum-Chemical Calculations of Antioxidant Activity

Herein, we used three analysis methods based on different reaction mechanisms, namely ORAC (HAT), FRAP (SET), and ABTS (SET or/and HAT), for an experimental assessment of antioxidant activity (AOA) of analogs of EDA 1aj, 4aj [67,68]. We noted that the values obtained with the ABTS and FRAP assays were significantly correlated (ρ = 0.702, p = 0.003), which was not the case for ABTS and ORAC (ρ = 0.391, p = 0.134) or FRAP and ORAC (ρ = 0.006, p = 0.985) (Spearman non-parametric correlation). These results may indicate the same SET mechanism of the antioxidant action of compounds 1a-j, 4aj in FRAP and ABTS tests, whereas another mechanism (HAT) is realized in ORAC assay.
Computational strategies are widely used to investigate AOA [69,70,71,72,73]. To identify parameters responsible for the antioxidant effect of the studied compounds 1aj and 4aj, we performed theoretical calculations of their electronic characteristics using the GAUSSIAN 09 program [74]. The most popular hybrid functional (B3LYP) [75] combined with the basis sets cc-pVDZ [76] was used throughout the work for geometry optimization and frequency calculations. Moreover, geometrical parameters of 4aj were optimized using the Polarizable Continuum Model (PCM) [77].
The scavenging activity of the aromatic compounds was shown to be directly proportional to the negative of the gaps computed as the energy difference between the HOMO and LUMO [78]; thus, we calculated such values for the synthesized analogs EDA 1aj and 4aj (Table 3).
As is known, the ionization potential IP is directly related to SET reactions; such an approach is valuable and provides primary physicochemical insight into the mechanism of action in tests where the SET mechanism is operational. Therefore, we analyzed the relationships of the AOA of compounds 1aj, 4aj defined in the ABTS and FRAP tests with their structure, using indices calculated from the vertical values of IP and EA. The use of only vertical values is due to the fact that the loss or capture of an electron by a molecule leads to an essential change in the structure and, consequently, the composition and energies of molecular orbitals. The vertical IP and EA quantities of all studied compounds 1aj, 4aj were collected in Table 3.
According to the experimental results, compounds 1 and 4 differed significantly in activity (Table 1), and the IP and gap calculation data confirmed this. For correct interpretation of the calculated quantum-chemical results, it is reasonable to consider the properties of compounds 1 and 4 within the following subgroups: polyfluoroalkyl (1ac and 4ac) and methyl (1d and 4d) derivatives with a phenyl substituent and polyfluoroalkyl (1ef and 4ei) and methyl (1j and 4j) derivatives with an NH fragment.
According to the data in Table 3, 4-aminopyrazolols 4aj have lower IP, gap, and η values compared to 4-hydroxyiminopyrazolones 1aj. This suggests that compounds 4aj should be more active than pyrazolones 1aj, and this was observed experimentally.
Compounds 1ad containing the N-Ph moiety have lower IP, gap, and η values compared to NH analogs 1ej. However, according to experimental data (Table 1), 4-hydroxyiminopyrazolones 1ad having a N-Ph group exhibited less antioxidant activity compared to HN-unsubstituted derivatives 1ej. Polyfluoroalkyl-containing compounds 1ac and 1ei have slightly higher IP values than methyl derivatives 1d and 1j. In this case, their gap and η values were nearly equal. According to the experiment, the fluorine-containing derivatives were somewhat less active than their methyl analogs.
4-Aminopyrazolol hydrochlorides 4 exhibited a pronounced antioxidant activity compared to 4-hydroxyiminopyrazolones 1 (Table 1), and this is consistent with lower IP, gap, and η values of compounds 4aj compared to 4-hydroxyiminopyrazolones 1aj. In the series of compounds 4, the property-changing tendency was similar to compounds 1 as a whole: compounds 4ad containing the N-phenyl moiety had lower IP, gap, and η values comparing to compounds 4ej. However, the activities of both sets 4ad and 4ej in the antioxidant experiment were approximately the same. Methyl derivatives 4d and 4j had a more pronounced antioxidant activity, and this was consistent with their lower IP, gap, and η values than those of fluorinated analogs 4a and 4e.
Next, we analyzed the dependencies of AOA (TEAC) on the calculated parameters of IP, gap, and η (eV) for compounds 1 and 4 according to the work of Horton et al. [78] for aromatic compounds having hydroxyl and amino groups.
Considering the possible dependencies of the experimentally found AOA values in the ABTS test (as more complete) on the calculated IP, gap, and η values for 4-hydroxyiminopyrazolones 1aj, we found that there was no direct dependence between the calculated IP and gap parameters with experimental AOA values.
At the same time, in the series of 4-aminopyrazolols 4 for all PhN-substituted compounds 4ad, there were good negative dependences between their AOA values and gap functions (Figure 2a). It was not possible to identify acceptable relationships for HN-unsubstituted analogs 4ej, because derivatives 4j and 4f containing Me- and H(CF2)2-substituents were expected to fall out of this series. For aminopyrazoles 4e,gi containing perfluoroalkyl substituent F(CF2)n, there was a dependence of their AOA on the IP function, with a favorable coefficient of determination (Figure 2b).
According to the lower gap values obtained for compounds 1ad and partially for aminopyrazoles 4ad with phenyl substituents, a higher AOA could be expected compared to HN-heterocycles 1ej and 4ej. However, this effect was observed for only one pair of 4-aminopyrazolols, 4c and 4i. Differences in theoretical and experimental data can be explained by the structure of these polyfunctional compounds and various transformations of molecules with a PhN substituent and HN group during antioxidant action. In addition, it is possible that the ABTS-binding effect of compounds 1 and 4 is realized not only by the SET mechanism but also partially by the HAT mechanism. This is supported by an incomplete correlation of the values from the ABTS test with the antioxidant action values from the FRAP test, where only the SET mechanism is implemented (Table 1).
In order to interpret the radical scavenging activity of antioxidants in reactions by the HAT mechanism, a common characteristic is known to be the bond-dissociation energy (BDE) [79,80]. Thus, using quantum-chemical calculations (B3LYP/cc-pVDZ method) in the approximation of the gas phase, we estimated the BDE (kJ/mol) values in the formation of possible radicals from compounds 1 and 4.
For 2-hydroxyiminopyrazolones 1aj, various radicals (examples for compounds 1a,d,e,j are shown in Scheme 2) are likely to be formed. It is possible to form only one radical, A1, from 2-hydroxyiminopyrazolone 1a, while for each of compounds 1d,e, two different radicals, B1, B2 and C1, C2, respectively, can be formed. In the case of NH-unsubstituted methylpyrazolone 1j, four D1D4 radicals can be generated. According to the calculated data for all compounds 1a,d,e,j, the formation of O-radicals A1, B1, C1, D1 through breaking of the -O-H bond of the hydroxyimine substituent is the most likely. The difference in the BDE values in the formation of O-radical and the other variants is presented in Scheme 2.
For 4-aminopyrazolol hydrochlorides 4aj, similar transformations can also be assumed, but taking into account the reactions of the hydroxyl and the amino groups, this leads to more complex and varied conversions. Thus, CF3-containing 4-aminopyrazolols 4a,e can theoretically form two J1, J2 (for 4a) or three K1K3 (for 4e) radicals (Scheme 3). Non-fluorinated analogs 4d,j can undergo reactions involving a Me group, thereby forming three L1L3 (for 4d) or four M1M4 (for 4j) radicals. Evaluating the difference between the calculated BDE values, the most likely process for compounds 4a,d,e,j was found to generate radicals J1, K1, L1, and M1, forming via -O-H bond breaking (the ΔBDE values are indicated in blue in Scheme 3) despite the position of this substituent in these pyrazoles changes. This process is similar to the transformations of structures 1a,d,e,j.
Given that by using quantum-chemical calculations for pyrazoles 1d,j, 4d,j, it was shown that the formation of O-radicals is energetically most advantageous in all cases (Scheme 2 and Scheme 3), we calculated the BDE values for -OH bond breaking in compounds 1aj and 4aj (Table 3). This revealed a linear relationship with a highly favorable coefficient of determination between the AOA from the ORAC test and the BDE (OH) values for a series of active NH-unsubstituted 4-hydroxyiminopyrazolones 1ej (Figure 3a).
All tested aminopyrazolols 4ae,gj were active in the ORAC test, and an acceptable linear relationship between AOA and BDE for polyfluoroalkyl derivatives of 4ac,e,gi, was found (Figure 3b), regardless of the presence of a Ph substituent at the nitrogen atom in the molecule. It was impossible to find such a pattern for non-fluorinated analogs owing to the limited samples including only 4d, 4j.
According to the calculated BDE values (Table 3), HN-unsubstituted 4-hydroxyiminopyrazolones 1ej are more active than 4-aminopyrazolols 4ej, and this is consistent with the AOA values from the ORAC test. According to the BDE values, the PhN-containing 4-aminopyrazolols 4ad should be more active than the HN analogs 4ej, which also coincides with the experiment.

2.5. Cytotoxicity Studies

To investigate the effect of the lead compound 4d as well as its NH analog 4j in comparison with EDA on cell viability, cytotoxicity studies for these compounds were performed on normal human dermal fibroblast (NHDF) cultures using the MTT/formazan assay [81].
EDA appeared to have a dose-dependent effect on NHDF culture viability, the inhibitory capability of which increased with rising concentration (Table 4).
Meanwhile, at a concentration of 100 µM, neither of the compounds 4d,j showed cytotoxicity against NHDF cell culture.

3. Materials and Methods

3.1. Chemistry

Acetic acid, chloroform, dimethylsulfoxide, ethanol, n-hexane, hydrochloric acid, and zinc (dust) were obtained from AO “VEKTON” (St. Petersburg, Russia). Pd/C (5 wt.%) was purchased from Alfa Aesar by Thermo Fisher Scientific (Kandel, Germany). The deuterosolvent DMSO-d6 was acquired from “SOLVEX” Limited Liability Company (Skolkovo Innovation Center, Moscow, Russia). Melting points were measured in open capillaries on a Stuart SMP30 melting point apparatus (Bibby Scientific Limited, Staffordshire, UK) and were uncorrected. The IR spectra were recorded on a Perkin Elmer Spectrum Two spectrophotometer (PerkinElmer, Waltham, MA, USA) using a frustrated total internal reflection accessory with a diamond crystal. The 1H and 19F NMR spectra were registered on a Bruker DRX-400 (400 or 376 MHz, respectively) or a Bruker AvanceIII 500 (500 or 470 MHz, respectively) (Bruker, Karlsruhe, Germany). The 13C NMR spectra were recorded on a Bruker AvanceIII 500 (125 MHz). The internal standard was SiMe4 (for 1H and 13C NMR spectra) and C6F6 (for 19F NMR spectra). The microanalyses (C, H, N) were carried out on a PerkinElmer PE 2400 series II elemental analyzer (PerkinElmer, Waltham, MA, USA). Chlorine was measured by mercurimetric titration. The column chromatography was performed on silica gel 60 (0.062–0.2 mm) (Macherey-Nagel GmbH & Co KG, Duren, Germany).
The initial 4-hydroxyiminopyrazol-5-ones 1ah were synthesized by the previously published method [52].

3.1.1. Synthesis of Rubazonic Acids 2a,b (General Procedure)

Pyrazolone 1a,b (5 mmol) was dissolved in glacial acetic acid (5 mL), and zinc dust (0.98 g, 15 mmol) was added. The reaction mixture was stirred for 4 h at room temperature (r.t.) and left standing overnight. Then, water (20 mL) was added, and the precipitate was filtered off and dried.
Bis[5-hydroxy-1-phenyl-3-(trifluoromethyl)-1H-pyrazol-4-yl]imine (2a). Yield 1.33 g (57%), red crystals, m.p. 188–189 °C. 1H NMR (500 MHz, CDCl3): δ 7.40–7.43, 7.50–7.53, 7.85–7.87 (all m, 10H, 2Ph); 17.16 (s, 1H, OH). 13C NMR (500 MHz, CDCl3): δ 119.23 (q, CF3, J 272.2 Hz); 121.29; 123.24; 128.33; 129.29; 136.42; 144.39 (q, C–CF3, J 37.4 Hz); 151.96. 19F NMR (500 MHz, CDCl3): δ 96.93 (s, CF3). Anal. calcd. for C20H11F6N5O2: C 51.40; H 2.37; N 14.99. C20H11F6N5O2. Found: C 51.54; H 2.32; N 14.94.
Crystallographic data for compound2a. The X-ray studies were performed on an “Xcalibur 3 CCD” diffractometer with a graphite monochromator, ω scanning with 1° step, λ(MoKα) 0.71073 Å radiation, T 295(2) K. Empirical absorption correction was applied. Using Olex2 [82], the structure was solved with the ShelXS [83] structure solution program using Direct Methods and refined with the ShelXL [83] refinement package using Least Squares minimization. All non-hydrogen atoms were refined in the anisotropic approximation; H-atoms at the C-H bonds were refined in the “rider” model with dependent displacement parameters. Empirical absorption correction was carried out through spherical harmonics, implemented in SCALE3 ABSPACK scaling algorithm by the program “CrysAlisPro 1.171.36.32” (Rigaku Oxford Diffraction, 2013).
Suitable red single crystals of compound 2a were obtained by slow crystallization from CHCl3. Main crystallographic data for 2a: C20H11F6N5O2, M = 467.34, space group C2/c, monoclinic, a 19.559(2), b 20.5241(19), c 9.6419(16) Å; β 93.650(12)°; V = 3862.8(8) Å3; Z 8; μ 0.147 mm−1; 303 refinement parameters; 11207 reflections measured; 3921 unique (Rint 0.0664), which were used in all calculations. The final wR2 was 0.1299 (all data), and R1 was 0.0588 [I >= 2σ (I)]. CCDC 2183294 contains the supplementary crystallographic data for this compound.
Bis[5-hydroxy-3-(pentafluoroethyl)-1-phenyl-1H-pyrazol-4-yl]imine (2b). Yield 1.28 g (45%), red crystals, m.p. 134–135 °C. 1H NMR (500 MHz, CDCl3): δ 7.40–7.43, 7.50–7.54, 7.86–7.88 (all m, 10H, 2Ph); 16.96 (s, 1H, OH). 19F NMR (CDCl3): δ 46.75 (br. s, 2F, CF2), 78.71 (br. s, 3F, CF3). Anal. calcd. for C22H11F10N5O2: C, 46.58; H, 1.95; N, 12.34. Found: C, 46.46; H, 2.04; N, 12.21.
N-(5-Hydroxy-3-(trifluoromethyl)-1-phenyl-1H-pyrazol-4-yl)acetamide (3a). Pyrazolone 1a (0.51 g, 2 mmol) was dissolved in glacial acetic acid (5 mL), and zinc dust (0.25 g, 4 mmol) was added. The reaction mixture was stirred for 30 min at r.t. Then, acetic anhydride (1 mL) was slowly added at 0 °C, and the reaction was left standing overnight at r.t. Then, water (20 mL) was added, and the precipitate was filtered off. The residue was purified by column chromatography (eluent–CHCl3/EtOH-10:1). Yield 0.43 g (75%), orange powder, m.p. 97–98 °C. IR: ν 3629, 3373, 3272, 3066 (OH, NH); 1649 (C=O); 1596, 1497, 1480 (C=N, C=C); 1228–1129 (C-F) cm−1. 1H NMR (400 MHz, CDCl3) δ 2.26 (s, 3H, Me), 7.27 (br. s, 1H, NH), 7.32–7.33, 7.44–7.48, 7.75–7.77 (all m, 5H, Ph); 11.45 (s, 1H, OH). 19F NMR (400 MHz, CDCl3) δ 100.45 (d, J 0.9 Hz, CF3). Anal. calcd. for C12H10F3N3O2: C, 50.53; H, 3.53; N, 14.73. Found: C, 50.44; H, 3.40; N, 14.58.

3.1.2. Synthesis of 4-Aminopyrazoles 4aj (General Procedure)

Method A. Tin (II) chloride dihydrate (8 mmol, 0.18 g) was dissolved in a minimum quantity of concentrated hydrochloric acid, and 4-hydroxyiminopyrazoles 1ac,fi (2 mmol) were added. The mixture was stirred for 2–3 h at r.t. or until a change in color of the initial compounds 1 was observed. The resulting solid was filtered off and washed by n-hexane to yield compounds 4ac,fi.
Method B. A solution of 4-hydroxyiminopyrazoles 1ce,j (1 mmol) in EtOH (10 mL) and 50 μL of HCl was hydrogenated in the presence of 5 wt.% Pd/C catalyst (10% mmol) in a steel autoclave under a hydrogen pressure of 5–7 bar and r.t. for 4–5 h. The solid impurities were filtered off, and the solvent was removed in vacuo. The resulting compound 4ce,j was washed with CHCl3.
4-Amino-3-(trifluoromethyl)-1-phenyl-1H-pyrazol-5-ol hydrochloride (4a). Yield 0.42 g (78%, method A), orange powder, m.p. 126–127 °C. IR: ν 3574, 3504, 3080, 2953 (OH, NH2); 1620, 1561, 1539, 1515 (C=N, C=C); 1153–1087 (CF) cm−1. 1H NMR (400 MHz, DMSO-d6): δ 7.42–7.46, 7.52–7.56, 7.69–7.71 (all m, 5H, Ph); NH2 and OH are overlapped by Ph protons. 1H NMR (The 1H NMR spectrum was registered along with 13C NMR after standing overnight [M1]) (500 MHz, DMSO-d6): a mixture of AH:AS (77:23): δ 7.21–7.24 (m, 1H, Ph AS); 7.23 (t, J 51.0 Hz, NH3+ AS); 7.43–7.47 (m, 1H+2H, Ph AH+AS); 7.53–7.57, 7.69–7.70 (both m, 4H, Ph AH); 7.93–7.94 (m, 2H, Ph AS); 9.42 (br.s, 3H, OH+NH2 AH). 13C NMR [M1] (500 MHz, DMSO-d6): a mixture of AH:AS: δ 118.53; 120.57 (q, J 270.4 Hz, CF3); 122.78 (br. m); 124.27; 124.73; 128.13 (br. m); 128.79; 129.25; 137.14 (q, J 35.1 Hz, C–CF3); 138.79; 147.06 (br. m); 152.52. 19F NMR (400 MHz, DMSO-d6): δ 102.35 (s, CF3). Anal. calcd. for C10H9ClF3N3O: C, 42.95; H, 3.24; N, 15.03; Cl, 12.68. Found: C, 42.94; H, 3.25; N, 15.02; Cl, 12.72.
4-Amino-3-(pentafluoroethyl)-1-phenyl-1H-pyrazol-5-ol hydrate hydrochloride (4b). Yield 0.52 g (75%, method A), light yellow powder, m.p. 176–178 °C. IR: ν 3057, 2884 (OH, NH2); 1616, 1529, 1506, 1486 (C=N, C=C); 1138–1106 (CF) cm−1. 1H NMR (500 MHz, DMSO-d6) δ 7.43–7.46, 7.54–7.57, 7.69–7.71 (all m, 5H, Ph); NH2 and OH are not observed due to deuterium exchange. 1H NMR [M1] (500 MHz, DMSO-d6): a mixture of AH:AS (40:60): δ 5.60 (br.s, 3H, OH+NH2 AH); 7.22–7.24 (m, 1H, Ph AS); 7.24 (t, J 51.0 Hz, NH3+ AS); 7.46–7.49 (m, 2H, Ph AS); 7.49–7.43 (m, 1H, Ph AH); 7.55–7.58, 7.68–7.69 (both m, 4H, Ph AH); 7.93–7.95 (m, 2H, Ph AS). 13C NMR [M1] (500 MHz, DMSO-d6): a mixture of AH:AS: δ 118.33–113.00; 116.80–119.82 (both m, C2F5); 118.47; 121.79; 122.50 (br. m); 124.80; 125.02; 127.02; 127.97 (br. m); 128.81; 129.02; 129.25; 131.59–132.18 (m); 135.93 (t, J 26.1 Hz, C–C2F5); 137.10 (br. m); 138.11; 138.73; 147.09 (br. m); 152.34. 19F NMR (500 MHz, DMSO-d6): δ 51.96 (m, 2F, CF2); 80.07 (t, J 2.6 Hz, 3F, CF3). Anal. calcd. for C11H11ClF5N3O2: C, 38.00; H, 3.19; N, 12.09; Cl, 10.20. Found: C, 37.79; H, 2.98; N, 11.98; Cl, 10.41.
5-Hydroxy-1-phenyl-3-(nonafluorobuthyl)-1H-pyrazol-4-ammonium hydrochloride (4c). Yield 0.56 g (65%, method A), 0.39 g (45%, method B), orange powder, m.p. 174–175 °C. IR: ν 3235, 3143, 3049, 2823 (OH, NH2); 1586, 1515, 1493, 1460 (C=N, C=C); 1133–1228 (CF) cm−1. 1H NMR (500MHz, DMSO-d6): δ 7.15 (t, 3H, NH3+, J 51.0 Hz); 7.21–7.25, 7.45–7.48, 7.93–7.94 (all m, 5H, Ph); OH is not observed due to deuterium exchange. 19F NMR (500MHz, DMSO-d6) δ 36.79–36.84, 40.34–40.40, 51.36–51.40 (all m, 6F, 3 CF2); 81.95 (t, J 8.5 Hz, 3F, CF3). Anal. calcd. for C13H9ClF9N3O: C, 36.34; H, 2.11; N, 9.78; Cl, 8.25. Found: C, 36.13; H, 1.99; N, 9.53; Cl, 8.46.
4-Amino-3-methyl-1-phenyl-1H-pyrazol-5-ol hydrochloride hydrate (4d). Yield 0.37 g (77%, method B), orange powder, m.p. 171 °C dec. IR: ν 3441, 3381, 3068, 2839 (OH, NH2); 1638, 1592, 1540, 1487 (C=N, C=C). 1H NMR (500 MHz, DMSO-d6): δ 2.26 (s, 3H, Me); 7.27–7.30, 7.46–7.50, 7.70–7.71 (all m, 5H, Ph); 10.15 (br.s, 3H, NH2 and OH). 13C NMR (500 MHz, DMSO-d6): δ 11.24; 96.81; 120.20; 125.82; 129.07 (2C); 137.29; 142.87. Anal. calcd. for C10H14ClN3O2: C, 49.29; H, 5.79; N, 17.24; Cl, 14.55. Found: C, 49.03; H, 5.74; N, 17.26, Cl, 14.68.
4-Amino-3-(trifluoromethyl)-1H-pyrazol-5-ol hydrochloride (4e). Yield 0.31 g (78%, method B), orange powder, m.p. 200–201 °C (lit. [35] m.p. 194–195 °C). IR: ν 3124, 3055, 2872 (OH, NH2); 1586, 1567, 1544, 1505 (C=N, C=C); 1147–1085 (CF) cm−1. 1H NMR (400 MHz, DMSO-d6): δ 10.43, 13.39 (both br.s, 4H, OH and NH2). 13C NMR (500 MHz, DMSO-d6): δ 92.76; 120.55 (q, J 268.9 Hz, CF3); 133.01 (m, C—CF3); 148.84. 19F NMR (400 MHz, DMSO-d6): δ 102.66 (s, CF3). Anal. calcd. for C4H5ClF3N3O: C, 23.60; H, 2.48; N, 20.64; Cl 17.42. Found: C, 23.37; H, 2.44; N, 20.19; Cl 17.56.
4-Amino-3-(1,1,2,2-tetrafluoroethyl)-1H-pyrazol-5-ol hydrate hydrochloride (4f). Yield 0.38 g (75%, method A), white powder, m.p. 144 °C dec. IR: ν 3198, 2963, 2802, 2680 (OH, NH2); 1592, 1571, 1523, 1511 (C=N, C=C); 1152–1089 (CF) cm−1. 1H NMR (400 MHz, DMSO-d6): δ 6.84 (t, J 51.4 Hz, 1H, HCF2); 10.12 (br.s, 3H, NH2 and OH), 13.34 (br.s, 1H, NH). 19F NMR (400 MHz, DMSO-d6): δ 24.90 (m, 2F, CF2); 49.78 (m, 2F, CF2). Anal. calcd. for C5H8ClF4N3O2: C, 23.68; H, 3.18; N, 16.57; Cl, 13.98. Found: C, 23.55; H, 2.98; N, 16.31; Cl, 14.05.
4-Amino-3-(pentafluoroethyl)-1H-pyrazol-5-ol hydrate hydrochloride (4g). Yield 0.36 g (66%, method A), white powder, m.p. 160 °C dec. IR: ν 3133, 2956, 2798, 2697 (OH, NH2); 1589, 1570, 1517, 1502 (C=N, C=C); 1147–1103 (CF) cm−1. 1H NMR (500 MHz, DMSO-d6): δ 10.56 (br.s, 3H, NH2 and OH); 13.56 (br.s, 1H, NH). 19F NMR (500 MHz, DMSO-d6): δ 52.30 (m, 2F, CF2); 79.90 (m, 3F, CF3). Anal. calcd. for C5H7ClF5N3O2: C, 22.11; H, 2.60; N, 15.47; Cl, 13.05. Found: C, 21.78; H, 2.25; N, 15.11; Cl, 13.42.
4-Amino-3-(heptafluoropropyl)-1H-pyrazol-5-ol hydrate hydrochloride (4h). Yield 0.46 g (72%, method A), white powder, m.p. 146 °C subl. IR: 3103, 3007, 2942, 2806 (OH, NH2); 1618, 1590, 1571 (C=N, C=C); 1149–1087 (C-F). 1H NMR (400 MHz, DMSO-d6): δ 9.72 (br. s, 3H, NH2 and OH); 13.54 (br.s, 1H, NH). 19F NMR (400 MHz, DMSO-d6): δ 36.89 (m, 2F, β-CF2), 53.80 (m, 2F, α-CF2), 83.06 (t, 3F, CF3, J 9.0 Hz). Anal. calcd. for C6H7ClF7N3O2: C, 22.41; H, 2.19; N, 13.07; Cl, 11.02. Found: C, 22.22; H, 2.07; N, 12.96; Cl, 11.25.
4-Amino-3-(nonafluorobutyl)-1H-pyrazol-5-ol hydrate hydrochloride (4i). Yield 0.50 g (68%, method A), white powder, m.p. 188–190 °C subl. IR: ν 3512, 3104, 2873, 2615 (OH, NH2); 1589, 1500 (C=N, C=C); 1230–1101 (CF) cm−1. 1H NMR (DMSO-d6): δ 10.11 (br.s, 1H, NH). 19F NMR (DMSO-d6): δ 37.48 (m, 2F, γ-CF2); 40.71 (d.d, J 17.1, 8.2 Hz, 2F, β-CF2); 54.61 (m, 2F, α-CF2); 82.11 (t, J 9.4 Hz, 3F, CF3). Anal. calcd. for C7H7ClF9N3O2: C, 22.63; H, 1.90; N, 11.31; Cl, 9.54. Found: C, 22.68; H, 1.85; N, 11.11; Cl, 9.39.
4-Amino-3-methyl-1H-pyrazol-5-ol hydrochloride (4j). Yield 0.20 g (67%, method B), orange powder, m.p. 200 °C dec (lit. [84] m.p. 200–203 °C). IR: 3222, 2877, 2694, 2664 (OH, NH2), 1674, 1639, 1590, 1579, 1556, 1532 (C=N, C=C). 1H NMR (500 MHz, DMSO-d6): δ 2.18 (s, 3H, Me); 9.84, 11.06 (both br.s, 3H, NH2 and OH). 13C NMR (500 MHz, DMSO-d6): δ 9.55; 95.47; 134.06; 154.34. Anal. calcd. for C4H8ClN3O: C, 32.12; H, 5.39; N, 28.09; Cl, 23.70. Found: C, 31.94; H, 5.38; N, 28.29; Cl, 23.90.
Copies of NMR spectra of compounds are available in Supplementary Material.

3.2. Biochemical Methods

3.2.1. ABTS Radical Cation Scavenging Activity Assay

Radical scavenging activity of the compounds was assessed using the ABTS radical cation (ABTS•+) decolorization assay [56] with some modifications [57]. ABTS (2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid)) was purchased from Tokyo Chemical Industry Co., Ltd. (Tokyo, Japan), and potassium persulfate (dipotassium peroxodisulfate), Trolox (6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid), and HPLC-grade ethanol and DMSO were obtained from Sigma-Aldrich (Saint Louis, MO, USA). Aqueous solutions were prepared using deionized water. All test compounds were dissolved in DMSO.
The solution of cation radical ABTS•+ was produced by incubation of an aqueous solution of 7 mM ABTS and 2.45 mM potassium persulfate solution in equal quantities for 12–16 h at room temperature in the dark. Radical scavenging capacity of the compounds was analyzed by mixing 10 μL of compound with 240 μL of ABTS•+ working solution in ethanol (100 μM final concentration). Data were given for 1 h of incubation of compounds with ABTS•+. The reduction in absorbance was measured spectrophotometrically at 734 nm using an xMark microplate UV/VIS microplate spectrophotometer (Bio-Rad, Hercules, CA, USA).
Ethanol blanks were run in each assay. Values were obtained from three replicates of each sample and three independent experiments.
Standard antioxidant Trolox was used as a reference compound. Antioxidant capacity as Trolox equivalent (TEAC) values was determined as the ratio between the slopes obtained from the linear correlation of the ABTS radical absorbance with the concentrations of tested compounds and Trolox. Quercetin was used as a positive control. For the test compounds that reduced ABTS•+ absorbance by more than 60% at 100 µM, we also determined the IC50 values (the compound concentration required for a 50% reduction of the ABTS radical). The compounds were tested in the concentration range of 0.5–400 µM.

3.2.2. The FRAP (Ferric Reducing Antioxidant Power) Assay

The FRAP (ferric reducing antioxidant power) assay measures the ability of antioxidants to reduce the ferric 2,4,6-tripyridyl-s-triazine complex [Fe (TPTZ)2]3+ to the intensely blue ferrous complex [Fe (TPTZ)2]2+ (λ = 593 nm) in acidic medium [58,59]. 2,4,6-Tris(pyridin-2-yl)-1,3,5-triazine (TPTZ), FeCl3·6H2O, Trolox, quercetin and DMSO were obtained from Sigma-Aldrich Chemical Co. (St. Louis, MO, USA). The FRAP reagent contained 2.5 mL of 10 mM TPTZ solution in 40 mM HCl, 2.5 mL of 20 mM FeCl3·6H2O water solution, and 25 mL of 0.3 M acetate buffer (pH 3.6). Aliquots of 10 µL of the test compound solution in DMSO (0.5 mM) were mixed with 240 µL of FRAP reagent, and the absorbance of the mixture was measured spectrophotometrically (FLUOStar OPTIMA microplate spectrophotometer (BMG Labtech, Ortenberg, Germany)) at 600 nm after 1 h incubation at 37 °C against a blank. Trolox was used as a reference compound. Quercetin was used as a positive control. A calibration curve with linear formula y = 0.03x (R2 = 0.998, p < 0.0001) was prepared for Trolox solutions in DMSO at a range of 2.5–100 µM, and the results are expressed as Trolox equivalents (TE), the values calculated as the ratio of the concentrations of Trolox and the test compound, resulting in the same effect.

3.2.3. ORAC Assay

AAPH (2,2’-azobis(2-amidinopropane) dihydrochloride), Trolox, and fluorescein sodium salt were purchased from Sigma-Aldrich, USA. PBS (phosphate-buffered saline) was from VWR Life Science. DMSO (dimethylsulfoxide) was obtained from Vekton, Russia.
The oxygen radical absorbance capacity (ORAC) was determined by the ability of compounds to protect fluorescein (FL) from peroxyl radicals that are generated by AAPH according to the known method [60,85] with minor variations. The reaction was carried out at 37 °C in black 96-well plates (SPL life sciences), and the final volume of the reaction mixture was 200 µL per well. The fluorescence was recorded by Tecan M1000 PRO microplate reader (Switzerland) at 485 nm excitation and 520 nm emission.
The 0.7 mM fluorescein sodium salt stock solution in PBS pH 7.4 (137 mM sodium chloride, 2.7 mM potassium chloride, and 10 mM phosphate buffer) was prepared and stored at 4–5 °C no longer than one month. An aliquot of the fluorescein stock solution was further diluted with PBS immediately before each assay to yield a 70 nM working solution. The 133 mM AAPH solution in PBS was prepared immediately before use. The test compounds and Trolox standard were dissolved in DMSO to 400 μM. The final concentration was 10 μM for all of the samples. For IC50 determination, a set of compound concentrations in the range of 0.5–50 μM was used.
The blank was composed of 15 μL PBS pH 7.4, 5 μL DMSO (final concentration: 2.5% (v/v)), 120 μL FL, and 60 μL AAPH, and it was included in each assay. PBS pH 7.4 (15 μL), antioxidant (the test compound or Trolox, 5 μL), and FL (120 μL, final concentration: 42 nM) solutions were placed in a black 96-well microplate and were pre-incubated for 15 min at 37 °C. AAPH solution (60 μL, final concentration 40 mM) was then added rapidly using a multichannel pipette. The fluorescence was recorded every minute for 2 h. A Trolox standard curve was also obtained in each assay. All reactions were carried out in quadruplicate, and two to four independent assays were performed for each sample.
Antioxidant curves (fluorescence vs. time) were first normalized to the curve of the blank (without antioxidant) corresponding to the same assay, and the net AUC corresponding to a sample was calculated by subtracting the AUC corresponding to the blank. The ORAC value was obtained by dividing the latter curve by the Trolox curve obtained in the same assay at 10 μM concentrations of compounds and Trolox. Final ORAC values, Trolox equivalents (TE), were expressed as μmol test compounds per μmol Trolox, where the value of TE for Trolox was taken as 1. Data were expressed as means ± SEM.

3.2.4. Inhibition In Vitro of Human Erythrocyte AChE, Equine Serum BChE, and Porcine Liver CES

Human erythrocyte acetylcholinesterase (AChE, EC 3.1.1.7), equine serum butyrylcholinesterase (BChE, EC 3.1.1.8), porcine liver carboxylesterase (CES, EC 3.1.1.1), acetylthiocholine iodide (ATCh), butyrylthiocholine iodide (BTCh), 4-nitrophenol acetate (4-NPA), and 5,5′-dithio-bis-(2-nitrobenzoic acid) (DTNB) were purchased from Sigma-Aldrich (Saint Louis, MO, USA).
All the kinetic experiments were performed under standard conditions, according to the protocol of IPAC RAS for a reversible inhibitors study.
Ellman’s colorimetric assay was used to measure AChE and BChE activity in 0.1 M K/Na phosphate buffer at pH 7.5, 25 °C [86]. Final concentrations of reactants were 0.33 mM DTNB, 0.02 unit/mL of AChE or BChE, and 1 mM of substrate (ATCh or BTCh, respectively). Reagent blanks consisted of reaction mixtures without enzyme to assess non-enzymatic hydrolysis of substrates. Porcine liver CES activity was assessed colorimetrically in 0.1 M K/Na phosphate buffer with pH 8.0 at 25 °C, by measuring the absorbance of 4-nitrophenol at 405 nm [87]. Final enzyme and substrate (4-NPA) concentrations were 0.02 unit/mL and 1 mM, respectively. Reagent blanks included all constituents except enzyme.
Test compounds were dissolved in DMSO. Reaction mixtures contained a final DMSO concentration of 2% (v/v). Enzyme inhibition was first assessed at a single concentration of 20 µM for each compound after a 10 min incubation at 25 °C in three separate experiments. Compounds inhibiting the enzyme by more than 30% were then selected for determination of the IC50 (inhibitor concentration resulting in 50% inhibition of control enzyme activity). Compounds (eight concentrations ranging between 10−11 and 10−4 M) were selected to achieve 20 to 80% inhibition and were incubated with each enzyme for 10 min at 25 °C. Substrate was then added, and residual enzyme activity relative to an inhibitor-free control was measured using a FLUOStar Optima microplate reader (LabTech, Ortenberg, Germany).

3.2.5. Cytotoxicity Studies

Cytotoxicity of compounds was evaluated using a culture of human dermal fibroblasts, which were isolated in the Laboratory of Cell Cultures of the Institute of Medical Cell Technologies, Ekaterinburg, Russia. Cells were seeded in 96-well plates in the inoculum dose of 2 × 105 cells/mL and cultured for 24 h in Dulbecco’s Modified Eagle Medium (DMEM), with 1% (w/v) glutamine in the presence of 10% (v/v) fetal bovine serum and gentamicin (50 mg/L) at 37 °C, with a humidified atmosphere of 5% (v/v) CO2. Then, compounds, which were solved in DMSO, were added to the wells at the final concentrations of 0.1, 1, 10, and 100 µM. The cells were incubated with compounds for 72 h, after which cell viability was assessed using the standard MTT assay [81]. The assay was carried out in four replicates with negative (culture medium) and positive (solution of the cytotoxic drug camptothecin at a concentration of 3 mM) controls, and the solvent control (DMSO). The results of the MTT assay were evaluated on a Tecan Infinite M200 PRO (Tecan Austria GmbH, Grödig, Austria) plate spectrophotometer by comparing the optical density of a formazan solution at 570 nm in the assay and control wells. The MTT staining of control cells was taken as 100%.

3.3. Quantum-Chemical Calculations

Adiabatic and vertical indices (electron affinity EA, ionization potential IP, and derived values) for 4-hydroxyiminopyrazolones 1aj were evaluated in the gas phase approximation (T 298 K and a pressure of 1 atm). The electronic parameters of the salts of 4-aminopyrazololes 4aj were calculated in the presence of the solvent (aqueous ethanol), because the effect of the solvent on cations is more pronounced than on neutral molecules. To obtain vertical values of electronic parameters, single-point calculations of anion and cation radicals were performed in the equilibrium geometry of the most stable conformer of a neutral species. In the case of salts 4aj, the cation structures were fully optimized, whereas single-point calculations were made for neutral or doubly charged species to estimate the EA or IP quantities, respectively. These indices for the studied compounds were calculated according to [88] using the following equations:
η = 1 2 I P E A
where η = the chemical hardness.
The ionization potential was estimated as the difference between the energies for the closed-shell N-electron and the open-shell (N − 1)-electron species [89]:
I P = E A + E A
The electron affinity of a molecule was determined in a similar manner using the open-shell (N + 1)-electron species:
E A = E A E A
where E is the total energy Etot in the calculations of vertical values.

3.4. Statistical Analyses

Experimental data were expressed as means ± SEM (n = 3 independent experiments) using GraphPad Prism version 6.05 for Windows, GraphPad Software (San Diego, CA, USA). Linear regressions with coefficients of determination (R2) were determined using Origin 6.1 for Windows (OriginLab, Northampton, MA, USA). Data were tested for normality (Gaussian distribution) using the D’Agostino–Pearson, Anderson–Darling, Shapiro–Wilks, and Kolmogorov–Smirnov tests as implemented in GraphPad Prism 9.4.1 for Windows GraphPad software (San Diego, CA, USA). The data from the ORAC antioxidant test passed all four normality tests, while the data from the FRAP antioxidant tests failed all four normality tests, and the data from the ABTS antioxidant test failed the Anderson–Darling normality test. Therefore, determinations of correlation coefficients between the three pairs of antioxidant data were carried out using the non-parametric Spearman ρ rather than the parametric Pearson r, again using Prism 9.4.1 for Windows. Statistical significance was set at p < 0.05.

4. Conclusions

We have proposed effective approaches for the reduction of 4-hydroxyiminopyrazol-5-ones, which allowed us to obtain a series of new 4-aminopyrazol-5-ols as analogs of Edaravone. It was found that 4-aminopyrazol-5-ols can exist only as salts, while in free form they either decompose or transform into imino-bis-pyrazolols (rubazonic acids). The structure of the resulting aminopyrazolol salts was confirmed by IR and NMR spectroscopy along with elemental analysis.
The antioxidant action of 4-aminopyrazol-5-ol hydrochlorides and their precursors—4-hydroxyiminopyrazol-5-ones—was investigated in three test systems (ABTS, FRAP, and ORAC) in comparison with EDA, CF3-EDA, and Trolox. These studies revealed a pronounced antioxidant activity of aminopyrazolols in all tests. The lead compound was 4-amino-3-methyl-1-phenylpyrazol-5-ol hydrochloride (4d). Its activity in ABTS and FRAP tests was comparable to that of EDA, CF3-EDA, and Trolox, whereas in the ORAC test, it showed higher efficacy than the reference compounds. The similar character of the change in antioxidant activity values with the variation in the structure of the compounds in the ABTS and FRAP experiments allowed us to conclude that the transformations of the test compounds in these tests occurred by the SET mechanism, while in the ORAC test, the transformations took place via the HAT mechanism.
Using quantum-chemical calculations, in general, the salts of 4-aminopyrazolols 4 were found to be characterized by lower values of the calculated electronic characteristics [gap, ionization potential (IP), and chemical hardness (η)] [89] compared to 4-hydroxyiminopyrazol-5-ones 1, and these results are consistent with their higher antioxidant activities determined experimentally in the ABTS and FRAP tests. Calculations showed that the most energetically favorable pathway was the formation of radicals through HO bond breaking in both 4-hydroxyiminopyrazolones 1 and 4-aminopyrazolols 4. The transformations of 4-aminopyrazolols 4 bearing more groups that are prone to the formation of radicals were not as unambiguous and more dependent on structural fragments. Thus, the antioxidant activity of PhN-4-aminopyrazolols 4ad in the ABTS test was determined by gap functions, whereas, for HN-unsubstituted analogs, correlation with IP functions was found only for perfluoroalkyl derivatives 4e,hi. The antioxidant activity in the ORAC test directly depended on the OH bond dissociation energy for active 4-hydroxyiminopyrazolones 1ej and for all group of 4-aminopyrazolols 4aj.
The esterase profile results demonstrated the absence of anticholinesterase activity for all the compounds and a moderate inhibition of CES by some 4-aminopyrazolols 4, especially for compounds 4c and 4i with a nonafluorobutyl substituent. The CES result should be taken into consideration in the potential therapeutic use of these compounds in high doses. In addition, the lead compound 4d and its NH-analog 4j did not show cytotoxicity against normal human fibroblasts, in contrast to EDA.
Thus, the investigation showed that 4-aminopyrazol-5-ols are promising antioxidant agents with activity at the Edaravone level and above. In particular, the lead compound 4d displayed good antioxidant properties, indicating its promise for potential future use as a novel therapeutic drug candidate in the treatment of diseases associated with oxidative stress. Moreover, the new synthesized aminopyrazolols are of interest for further modification as antioxidant pharmacophores, e.g., for conjugation to an anticholinesterase fragment and for the creation of multifunctional drugs for the treatment of neurodegenerative diseases.

Supplementary Materials

The following information is available online at https://www.mdpi.com/article/10.3390/molecules27227722/s1: Figure S1: 1H NMR spectrum of bis[5-hydroxy-1-phenyl-3-(trifluoromethyl)-1H-pyrazol-4-yl]imine (2a); Figure S2: 13C NMR spectrum of bis[5-hydroxy-1-phenyl-3-(trifluoromethyl)-1H-pyrazol-4-yl]imine (2a); Figure S3: 19F NMR spectrum of bis[5-hydroxy-1-phenyl-3-(trifluoromethyl)-1H-pyrazol-4-yl]imine (2a); Figure S4: 1H NMR spectrum of bis[5-hydroxy-3-(pentafluoroethyl)-1-phenyl-1H-pyrazol-4-yl]imine (2b); Figure S5: 19F NMR spectrum of bis[5-hydroxy-3-(pentafluoroethyl)-1-phenyl-1H-pyrazol-4-yl]imine (2b); Figure S6: 1H NMR spectrum of N-(5-hydroxy-3-(trifluoromethyl)-1-phenyl-1H-pyrazol-4-yl)acetamide (3a); Figure S7: 19F NMR spectrum of N-(5-hydroxy-3-(trifluoromethyl)-1-phenyl-1H-pyrazol-4-yl)acetamide (3a); Figure S8: 1H NMR spectrum of 4-amino-3-trifluoromethyl-1-phenyl-1H-pyrazol-5-ol hydrochloride (4a); Figure S9: 1H NMR spectrum of 4-amino-3-trifluoromethyl-1-phenyl-1H-pyrazol-5-ol hydrochloride (4a) after standing overnight; Figure S10: 13C NMR spectrum of 4-amino-3-trifluoromethyl-1-phenyl-1H-pyrazol-5-ol hydrochloride (4a); Figure S11: 19F NMR spectrum of 4-amino-3-trifluoromethyl-1-phenyl-1H-pyrazol-5-ol hydrochloride (4a); Figure S12: 1H NMR spectrum of 4-amino-3-pentafluoroethyl-1-phenyl-1H-pyrazol-5-ol hydrate hydrochloride (4b); Figure S13: 1H NMR spectrum of 4-amino-3-pentafluoroethyl-1-phenyl-1H-pyrazol-5-ol hydrate hydrochloride (4b) after standing overnight; Figure S14: 13C NMR spectrum of 4-amino-3-pentafluoroethyl-1-phenyl-1H-pyrazol-5-ol hydrate hydrochloride (4b); Figure S15: 19F NMR spectrum of 4-amino-3-pentafluoroethyl-1-phenyl-1H-pyrazol-5-ol hydrate hydrochloride (4b); Figure S16: 1H NMR spectrum of 5-hydroxy-1-phenyl-3-nonafluorobuthyl-1H-pyrazol-4-ammonium hydrochloride (4c); Figure S17: 19F NMR spectrum of 5-hydroxy-1-phenyl-3-nonafluorobuthyl-1H-pyrazol-4-ammonium hydrochloride (4c); Figure S18: 1H NMR spectrum of 4-amino-3-methyl-1-phenyl-1H-pyrazol-5-ol hydrochloride hydrate (4d); Figure S19: 13C NMR spectrum of 4-amino-3-methyl-1-phenyl-1H-pyrazol-5-ol hydrochloride hydrate (4d); Figure S20: 1H NMR spectrum of 4-amino-3-trifluoromethyl-1H-pyrazol-5-ol hydrochloride (4e); Figure S21: 13C NMR spectrum of 4-amino-3-trifluoromethyl-1H-pyrazol-5-ol hydrochloride (4e); Figure S22: 19F NMR spectrum of 4-amino-3-trifluoromethyl-1H-pyrazol-5-ol hydrochloride (4e); Figure S23: 1H NMR spectrum of 4-amino-3-(1,1,2,2-tetrafluoroethyl)-1H-pyrazol-5-ol hydrate hydrochloride (4f); Figure S24: 19F NMR spectrum of 4-amino-3-(1,1,2,2-tetrafluoroethyl)-1H-pyrazol-5-ol hydrate hydrochloride (4f); Figure S25: 1H NMR spectrum of 4-amino-3-(pentafluoroethyl)-1H-pyrazol-5-ol hydrate hydrochloride (4g); Figure S26: 19F NMR spectrum of 4-amino-3-(pentafluoroethyl)-1H-pyrazol-5-ol hydrate hydrochloride (4g); Figure S27: 1H NMR spectrum of 4-amino-3-(heptafluoropropyl)-1H-pyrazol-5-ol hydrate hydrochloride (4h); Figure S28: 19F NMR spectrum of 4-amino-3-(heptafluoropropyl)-1H-pyrazol-5-ol hydrate hydrochloride (4h); Figure S29: 1H NMR spectrum of 4-amino-3-(nonafluorobutyl)-1H-pyrazol-5-ol hydrate hydrochloride (4i); Figure S30: 19F NMR spectrum of 4-amino-3-(nonafluorobutyl)-1H-pyrazol-5-ol hydrate hydrochloride (4i); Figure S31: 1H NMR spectrum of 4-amino-3-methyl-1H-pyrazol-5-ol hydrochloride (4j); Figure S32:13C NMR spectrum of 4-amino-3-methyl-1H-pyrazol-5-ol hydrochloride (4j).

Author Contributions

Conceptualization, methodology, Y.V.B., G.F.M., O.P.K., S.L.K., V.I.S.; investigation, N.A.A., E.V.S., G.A.T., D.A.G., S.S.B., N.V.K., E.V.R., N.P.B., O.G.S., M.V.U.; writing—original draft preparation, Y.V.B., E.V.S., S.S.B., O.P.K., N.V.K., E.V.R., N.P.B.; writing—review and editing, Y.V.B., E.V.S., G.F.M., R.J.R., N.V.K., E.V.R., N.P.B.; project administration, V.I.S., G.F.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation (Agreement No. 075-15-2020-777).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Analytical studies were carried out using equipment of the Center for Joint Use “Spectroscopy and Analysis of Organic Compounds” at the Postovsky Institute of Organic Synthesis of UB RAS. We thank the “Centre for Collective Use of IPAC RAS” (IPAC research topic FFSN-2021-0005) for use of equipment for biochemical studies cited in the Methods.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds 1a–j, 2a,b, 3a, 4a–j are available from the authors.

References

  1. Neha, K.; Haider, M.R.; Pathak, A.; Yar, M.S. Medicinal prospects of antioxidants: A review. Eur. J. Med. Chem. 2019, 178, 687–704. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, Z.-Q. Bridging free radical chemistry with drug discovery: A promising way for finding novel drugs efficiently. Eur. J. Med. Chem. 2020, 189, 112020. [Google Scholar] [CrossRef] [PubMed]
  3. Silva, V.L.M.; Elguero, J.; Silva, A.M.S. Current progress on antioxidants incorporating the pyrazole core. Eur. J. Med. Chem. 2018, 156, 394–429. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, Z.; Dai, X.; Li, C.; Wang, X.; Tian, J.; Feng, Y.; Xie, J.; Ma, C.; Nie, Z.; Fan, P.; et al. Pyrazolone structural motif in medicinal chemistry: Retrospect and prospect. Eur. J. Med. Chem. 2020, 186, 111893. [Google Scholar] [CrossRef]
  5. Queiroz, A.N.; Martins, C.C.; Santos, K.L.B.; Carvalho, E.S.; Owiti, A.O.; Oliveira, K.R.M.; Herculano, A.M.; da Silva, A.B.F.; Borges, R.S. Experimental and theoretical study on structure-tautomerism among edaravone, isoxazolone, and their heterocycles derivatives as antioxidants. Saudi Pharm. J. 2020, 28, 819–827. [Google Scholar] [CrossRef] [PubMed]
  6. Abe, K.; Yuki, S.; Kogure, K. Strong attenuation of ischemic and postischemic brain edema in rats by a novel free radical scavenger. Stroke 1988, 19, 480–485. [Google Scholar] [CrossRef] [Green Version]
  7. Miyaji, Y.; Yoshimura, S.; Sakai, N.; Yamagami, H.; Egashira, Y.; Shirakawa, M.; Uchida, K.; Kageyama, H.; Tomogane, Y. Effect of Edaravone on Favorable Outcome in Patients with Acute Cerebral Large Vessel Occlusion: Subanalysis of RESCUE-Japan Registry. Neurol. Med. Chir. 2015, 55, 241–247. [Google Scholar] [CrossRef] [Green Version]
  8. Sun, Z.; Xu, Q.; Gao, G.; Zhao, M.; Sun, C. Clinical observation in edaravone treatment for acute cerebral infarction. Niger. J. Clin. Pract. 2019, 22, 1324–1327. [Google Scholar] [CrossRef]
  9. Dash, R.P.; Babu, R.J.; Srinivas, N.R. Two Decades-Long Journey from Riluzole to Edaravone: Revisiting the Clinical Pharmacokinetics of the Only Two Amyotrophic Lateral Sclerosis Therapeutics. Clin. Pharmacokinet. 2018, 57, 1385–1398. [Google Scholar] [CrossRef] [PubMed]
  10. Luo, L.; Song, Z.; Li, X.; Huiwang; Zeng, Y.; Qinwang; Meiqi; He, J. Efficacy and safety of edaravone in treatment of amyotrophic lateral sclerosis—A systematic review and meta-analysis. Neurol. Sci. 2018, 40, 235–241. [Google Scholar] [CrossRef]
  11. Shefner, J.; Heiman-Patterson, T.; Pioro, E.P.; Wiedau-Pazos, M.; Liu, S.; Zhang, J.; Agnese, W.; Apple, S. Long-term edaravone efficacy in amyotrophic lateral sclerosis: Post-hoc analyses of Study 19 (MCI186-19). Muscle Nerve 2019, 61, 218–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bailly, C. Potential use of edaravone to reduce specific side effects of chemo-, radio- and immuno-therapy of cancers. Int. Immunopharmacol. 2019, 77, 105967. [Google Scholar] [CrossRef]
  13. Zhang, W.-W.; Bai, F.; Wang, J.; Zheng, R.-H.; Yang, L.-W.; James, E.; Zhao, Z.-Q. Edaravone inhibits pressure overload-induced cardiac fibrosis and dysfunction by reducing expression of angiotensin II AT1 receptor. Drug Des. Dev. Ther. 2017, 11, 3019–3033. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, Z.-Q.; Pang, X.-F.; Zhang, L.-H.; Bai, F.; Wang, N.-P.; McKallip, R.; Garner, R. Attenuation of myocardial fibrosis with curcumin is mediated by modulating expression of angiotensin II AT1/AT2 receptors and ACE2 in rats. Drug Des. Dev. Ther. 2015, 9, 6043–6054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kikuchi, K.; Takeshige, N.; Miura, N.; Morimoto, Y.; Ito, T.; Tancharoen, S.; Miyata, K.E.I.; Kikuchi, C.; Iida, N.; Uchikado, H.; et al. Beyond free radical scavenging: Beneficial effects of edaravone (Radicut) in various diseases (Review). Exp. Ther. Med. 2012, 3, 3–8. [Google Scholar] [CrossRef] [Green Version]
  16. Raymer, B.; Bhattacharya, S.K. Lead-like Drugs: A Perspective. J. Med. Chem. 2018, 61, 10375–10384. [Google Scholar] [CrossRef] [PubMed]
  17. Higashi, Y.; Jitsuiki, D.; Chayama, K.; Yoshizumi, M. Edaravone (3-Methyl-1-Phenyl-2-Pyrazolin-5-one), A Novel Free Radical Scavenger, for Treatment of Cardiovascular Diseases. Recent Pat. Cardiovasc. Drug Discov. 2006, 1, 85–93. [Google Scholar] [CrossRef] [PubMed]
  18. Ren, Y.; Wei, B.; Song, X.; An, N.; Zhou, Y.; Jin, X.; Zhang, Y. Edaravone’s free radical scavenging mechanisms of neuroprotection against cerebral ischemia: Review of the literature. Int. J. Neurosci. 2014, 125, 555–565. [Google Scholar] [CrossRef]
  19. Kamogawa, E.; Sueishi, Y. A multiple free-radical scavenging (MULTIS) study on the antioxidant capacity of a neuroprotective drug, edaravone as compared with uric acid, glutathione, and trolox. Bioorg. Med. Chem. Lett. 2014, 24, 1376–1379. [Google Scholar] [CrossRef]
  20. Yamamoto, Y.; Kuwahara, T.; Watanabe, K.; Watanabe, K. Antioxidant activity of 3-methyl-1-phenyl-2-pyrazolin-5-one. Redox Rep. 2016, 2, 333–338. [Google Scholar] [CrossRef]
  21. Homma, T.; Kobayashi, S.; Sato, H.; Fujii, J. Edaravone, a free radical scavenger, protects against ferroptotic cell death in vitro. Exp. Cell Res. 2019, 384, 111592. [Google Scholar] [CrossRef] [PubMed]
  22. Bailly, C.; Hecquet, P.-E.; Kouach, M.; Thuru, X.; Goossens, J.-F. Chemical reactivity and uses of 1-phenyl-3-methyl-5-pyrazolone (PMP), also known as edaravone. Bioorg. Med. Chem. 2020, 28, 115463. [Google Scholar] [CrossRef] [PubMed]
  23. Pérez-González, A.; Galano, A. OH Radical Scavenging Activity of Edaravone: Mechanism and Kinetics. J. Phys. Chem. B 2010, 115, 1306–1314. [Google Scholar] [CrossRef]
  24. Yamamoto, Y. Plasma marker of tissue oxidative damage and edaravone as a scavenger drug against peroxyl radicals and peroxynitrite. J. Clin. Biochem. Nutr. 2017, 60, 49–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Nakagawa, H.; Ohyama, R.; Kimata, A.; Suzuki, T.; Miyata, N. Hydroxyl radical scavenging by edaravone derivatives: Efficient scavenging by 3-methyl-1-(pyridin-2-yl)-5-pyrazolone with an intramolecular base. Bioorg. Med. Chem. Lett. 2006, 16, 5939–5942. [Google Scholar] [CrossRef] [PubMed]
  26. Chegaev, K.; Cena, C.; Giorgis, M.; Rolando, B.; Tosco, P.; Bertinaria, M.; Fruttero, R.; Carrupt, P.-A.; Gasco, A. Edaravone Derivatives Containing NO-Donor Functions. J. Med. Chem. 2008, 52, 574–578. [Google Scholar] [CrossRef]
  27. Rolando, B.; Filieri, A.; Chegaev, K.; Lazzarato, L.; Giorgis, M.; De Nardi, C.; Fruttero, R.; Martel, S.; Carrupt, P.-A.; Gasco, A. Synthesis physicochemical profile and PAMPA study of new NO-donor edaravone co-drugs. Bioorg. Med. Chem. 2012, 20, 841–850. [Google Scholar] [CrossRef]
  28. Gaffer, H.E.; Abdel-Fattah, S.; Etman, H.A.; Abdel-Latif, E. Synthesis and Antioxidant Activity of Some New Thiazolyl-Pyrazolone Derivatives. J. Heterocycl. Chem. 2017, 54, 331–340. [Google Scholar] [CrossRef]
  29. Polkam, N.; Ramaswamy, V.R.; Rayam, P.; Allaka, T.R.; Anantaraju, H.S.; Dharmarajan, S.; Perumal, Y.; Gandamalla, D.; Yellu, N.R.; Balasubramanian, S.; et al. Synthesis, molecular properties prediction and anticancer, antioxidant evaluation of new edaravone derivatives. Bioorg. Med. Chem. Lett. 2016, 26, 2562–2568. [Google Scholar] [CrossRef]
  30. Qiang, X.; Li, Y.; Yang, X.; Luo, L.; Xu, R.; Zheng, Y.; Cao, Z.; Tan, Z.; Deng, Y. DL-3-n-butylphthalide-Edaravone hybrids as novel dual inhibitors of amyloid- β aggregation and monoamine oxidases with high antioxidant potency for Alzheimer’s therapy. Bioorg. Med. Chem. Lett. 2017, 27, 718–722. [Google Scholar] [CrossRef]
  31. Purohit, M.K.; Chakka, S.K.; Scovell, I.; Neschadim, A.; Bello, A.M.; Salum, N.; Katsman, Y.; Bareau, M.C.; Branch, D.R.; Kotra, L.P. Structure–activity relationships of pyrazole derivatives as potential therapeutics for immune thrombocytopenias. Bioorg. Med. Chem. 2014, 22, 2739–2752. [Google Scholar] [CrossRef] [PubMed]
  32. Hussain, G.; Ather, M.; Khan, M.U.A.; Saeed, A.; Saleem, R.; Shabir, G.; Channar, P.A. Synthesis and characterization of chromium (III), iron (II), copper (II) complexes of 4-amino-1-(p-sulphophenyl)-3-methyl-5-pyrazolone based acid dyes and their applications on leather. Dyes Pigment. 2016, 130, 90–98. [Google Scholar] [CrossRef]
  33. Uchida, T.; Kawanishi, T.; Arita, E.; Sato, Y.; Aoki, S.; Ogawa, M.; Nakano, T.; Hasegawa, Y. Preparation of Oxidation Coloring Compounds, 1-(Sulfonated phenyl)-4-amino-5-pyrazolones. Patent JP 2012214419, 31 October 2011. Available online: https://patents.google.com/patent/JP2012214419A/en (accessed on 1 October 2022).
  34. Isaad, J. Highly water soluble dyes based on pyrazolone derivatives of carbohydrates. Tetrahedron 2013, 69, 2239–2250. [Google Scholar] [CrossRef]
  35. Drummond, J.; Johnson, G.; Nickell, D.G.; Ortwine, D.F.; Bruns, R.F.; Welbaum, B. Evaluation and synthesis of amino-hydroxy isoxazoles and pyrazoles as potential glycine agonists. J. Med. Chem. 1998, 32, 2116–2128. [Google Scholar] [CrossRef]
  36. Kawase, M.; Koiwai, H.; Yamano, A.; Miyamae, H. Regioselective reaction of mesoionic 4-trifluoroacetyl-1,3-oxazolium-5-olates and phenylhydrazine: Synthesis of trifluoromethyl substituted pyrazole and 1,2,4-triazine derivatives. Tetrahedron Lett. 1998, 39, 663–666. [Google Scholar] [CrossRef]
  37. Kaur, K.; Kumar, V.; Gupta, G.K. Trifluoromethylpyrazoles as anti-inflammatory and antibacterial agents: A review. J. Fluor. Chem. 2015, 178, 306–326. [Google Scholar] [CrossRef]
  38. Sloop, J.C.; Holder, C.; Henary, M. Selective Incorporation of Fluorine in Pyrazoles. Eur. J. Org. Chem. 2015, 2015, 3405–3422. [Google Scholar] [CrossRef]
  39. O’Hagan, D. Fluorine in health care: Organofluorine containing blockbuster drugs. J. Fluorine Chem. 2010, 131, 1071–1081. [Google Scholar] [CrossRef]
  40. Hagmann, W.K. The Many Roles for Fluorine in Medicinal Chemistry. J. Med. Chem. 2008, 51, 4359–4369. [Google Scholar] [CrossRef]
  41. Inoue, M.; Sumii, Y.; Shibata, N. Contribution of Organofluorine Compounds to Pharmaceuticals. ACS Omega 2020, 5, 10633–10640. [Google Scholar] [CrossRef]
  42. Mykhailiuk, P.K. Fluorinated Pyrazoles: From Synthesis to Applications. Chem. Rev. 2020, 121, 1670–1715. [Google Scholar] [CrossRef] [PubMed]
  43. Kamble, O.; Dandela, R.; Shinde, S. Recent Innovations of Organo-fluorine Synthesis and Pharmacokinetics. Curr. Org. Chem. 2021, 25, 2650–2665. [Google Scholar] [CrossRef]
  44. Politanskaya, L.V.; Selivanova, G.A.; Panteleeva, E.V.; Tretyakov, E.V.; Platonov, V.E.; Nikul’shin, P.V.; Vinogradov, A.S.; Zonov, Y.V.; Karpov, V.M.; Mezhenkova, T.V.; et al. Organofluorine chemistry: Promising growth areas and challenges. Russ. Chem. Rev. 2019, 88, 425–569. [Google Scholar] [CrossRef]
  45. Jeffries, B.; Wang, Z.; Graton, J.; Holland, S.D.; Brind, T.; Greenwood, R.D.R.; Le Questel, J.-Y.; Scott, J.S.; Chiarparin, E.; Linclau, B. Reducing the Lipophilicity of Perfluoroalkyl Groups by CF2–F/CF2–Me or CF3/CH3 Exchange. J. Med. Chem. 2018, 61, 10602–10618. [Google Scholar] [CrossRef] [Green Version]
  46. Mohd Faudzi, S.M.; Leong, S.W.; Auwal, F.A.; Abas, F.; Wai, L.K.; Ahmad, S.; Tham, C.L.; Shaari, K.; Lajis, N.H.; Yamin, B.M. In silico studies, nitric oxide, and cholinesterases inhibition activities of pyrazole and pyrazoline analogs of diarylpentanoids. Arch. Pharm. 2020, 354, e2000161. [Google Scholar] [CrossRef]
  47. Gutti, G.; Kumar, D.; Paliwal, P.; Ganeshpurkar, A.; Lahre, K.; Kumar, A.; Krishnamurthy, S.; Singh, S.K. Development of pyrazole and spiropyrazoline analogs as multifunctional agents for treatment of Alzheimer’s disease. Bioorg. Chem. 2019, 90, 103080. [Google Scholar] [CrossRef]
  48. Min, L.-J.; Zhai, Z.-W.; Shi, Y.-X.; Han, L.; Tan, C.-X.; Weng, J.-Q.; Li, B.-J.; Zhang, Y.-G.; Liu, X.-H. Synthesis and biological activity of acyl thiourea containing difluoromethyl pyrazole motif. Phosphorus Sulfur Silicon Relat. Elem. 2019, 195, 22–28. [Google Scholar] [CrossRef]
  49. Tok, F.; Koçyiğit-Kaymakçıoğlu, B.; Sağlık, B.N.; Levent, S.; Özkay, Y.; Kaplancıklı, Z.A. Synthesis and biological evaluation of new pyrazolone Schiff bases as monoamine oxidase and cholinesterase inhibitors. Bioorg. Chem. 2019, 84, 41–50. [Google Scholar] [CrossRef]
  50. Turkan, F.; Cetin, A.; Taslimi, P.; Gulçin, İ. Some pyrazoles derivatives: Potent carbonic anhydrase, α-glycosidase, and cholinesterase enzymes inhibitors. Arch. Pharm. 2018, 351, e1800200. [Google Scholar] [CrossRef]
  51. Bao, X.; Wei, S.; Qian, X.; Qu, J.; Wang, B.; Zou, L.; Ge, G. Asymmetric Construction of a Multi-Pharmacophore-Containing Dispirotriheterocyclic Scaffold and Identification of a Human Carboxylesterase 1 Inhibitor. Org. Lett. 2018, 20, 3394–3398. [Google Scholar] [CrossRef]
  52. Burgart, Y.V.; Agafonova, N.A.; Shchegolkov, E.V.; Maslova, V.V.; Triandafilova, G.A.; Solodnikov, S.Y.; Krasnykh, O.P.; Saloutin, V.I. New one-pot synthesis of 4-hydroxyimino-5-polyfluoroalkylpyrazol-3-ones, their structure and biological activity. Chem. Heterocycl. Compd. 2019, 55, 52–59. [Google Scholar] [CrossRef]
  53. Burgart, Y.V.; Agafonova, N.A.; Shchegolkov, E.V.; Krasnykh, O.P.; Kushch, S.O.; Evstigneeva, N.P.; Gerasimova, N.A.; Maslova, V.V.; Triandafilova, G.A.; Solodnikov, S.Y.; et al. Multiple biological active 4-aminopyrazoles containing trifluoromethyl and their 4-nitroso-precursors: Synthesis and evaluation. Eur. J. Med. Chem. 2020, 208, 112768. [Google Scholar] [CrossRef] [PubMed]
  54. Hänsel, W. 4-(5-Hydroxy-4-pyrazolylimino)-2-pyrazolin-5-one und ihre Metallchelate, I Synthese von 4-(5-Hydroxy-4-pyrazolylimino)-2-pyrazolin-5-onen (Rubazonsäuren) und strukturanalogen Verbindungen. Justus Liebigs Ann. Der Chem. 1976, 1976, 1380–1394. [Google Scholar] [CrossRef]
  55. Ueda, T.; Oda, N.; Ito, I. Studies on synthetic methods for 5-amino-4(3H)-pyrimidones. I. A novel ring expansion reaction of 4-aminoantipyrines to 5-amino-4(3H)-pyrimidones. Chem. Pharm. Bull. 1980, 28, 2144–2147. [Google Scholar] [CrossRef] [Green Version]
  56. Re, R.; Pellegrini, N.; Proteggente, A.; Pannala, A.; Yang, M.; Rice-Evans, C. Antioxidant activity applying an improved ABTS radical cation decolorization assay. Free Radic. Biol. Med. 1999, 26, 1231–1237. [Google Scholar] [CrossRef]
  57. Makhaeva, G.F.; Elkina, N.A.; Shchegolkov, E.V.; Boltneva, N.P.; Lushchekina, S.V.; Serebryakova, O.G.; Rudakova, E.V.; Kovaleva, N.V.; Radchenko, E.V.; Palyulin, V.A.; et al. Synthesis, molecular docking, and biological evaluation of 3-oxo-2-tolylhydrazinylidene-4,4,4-trifluorobutanoates bearing higher and natural alcohol moieties as new selective carboxylesterase inhibitors. Bioorg. Chem. 2019, 91, 103097. [Google Scholar] [CrossRef] [PubMed]
  58. Benzie, I.F.F.; Strain, J.J. [2] Ferric reducing/antioxidant power assay: Direct measure of total antioxidant activity of biological fluids and modified version for simultaneous measurement of total antioxidant power and ascorbic acid concentration. Methods Enzymol. 1999, 299, 15–27. [Google Scholar] [CrossRef]
  59. Makhaeva, G.F.; Kovaleva, N.V.; Rudakova, E.V.; Boltneva, N.P.; Lushchekina, S.V.; Faingold, I.I.; Poletaeva, D.A.; Soldatova, Y.V.; Kotelnikova, R.A.; Serkov, I.V.; et al. New Multifunctional Agents Based on Conjugates of 4-Amino-2,3-polymethylenequinoline and Butylated Hydroxytoluene for Alzheimer’s Disease Treatment. Molecules 2020, 25, 5891. [Google Scholar] [CrossRef]
  60. Ou, B.; Hampsch-Woodill, M.; Prior, R.L. Development and Validation of an Improved Oxygen Radical Absorbance Capacity Assay Using Fluorescein as the Fluorescent Probe. J. Agric. Food. Chem. 2001, 49, 4619–4626. [Google Scholar] [CrossRef]
  61. Gotsbacher, M.P.; Telfer, T.J.; Witting, P.K.; Double, K.L.; Finkelstein, D.I.; Codd, R. Analogues of desferrioxamine B designed to attenuate iron-mediated neurodegeneration: Synthesis, characterisation and activity in the MPTP-mouse model of Parkinson’s disease. Metallomics 2017, 9, 852–864. [Google Scholar] [CrossRef]
  62. Kraus, R.L.; Pasieczny, R.; Lariosa-Willingham, K.; Turner, M.S.; Jiang, A.; Trauger, J.W. Antioxidant properties of minocycline: Neuroprotection in an oxidative stress assay and direct radical-scavenging activity. J. Neurochem. 2005, 94, 819–827. [Google Scholar] [CrossRef] [PubMed]
  63. Makhaeva, G.F.; Rudakova, E.V.; Serebryakova, O.G.; Aksinenko, A.Y.; Lushchekina, S.V.; Bachurin, S.O.; Richardson, R.J. Esterase profiles of organophosphorus compounds in vitro predict their behavior in vivo. Chem. Biol. Interact. 2016, 259, 332–342. [Google Scholar] [CrossRef] [PubMed]
  64. Makhaeva, G.F.; Radchenko, E.V.; Palyulin, V.A.; Rudakova, E.V.; Aksinenko, A.Y.; Sokolov, V.B.; Zefirov, N.S.; Richardson, R.J. Organophosphorus compound esterase profiles as predictors of therapeutic and toxic effects. Chem. Biol. Interact. 2013, 203, 231–237. [Google Scholar] [CrossRef] [PubMed]
  65. Makhaeva, G.F.; Lushchekina, S.V.; Boltneva, N.P.; Serebryakova, O.G.; Kovaleva, N.V.; Rudakova, E.V.; Elkina, N.A.; Shchegolkov, E.V.; Burgart, Y.V.; Stupina, T.S.; et al. Novel potent bifunctional carboxylesterase inhibitors based on a polyfluoroalkyl-2-imino-1,3-dione scaffold. Eur. J. Med. Chem. 2021, 218, 113385. [Google Scholar] [CrossRef] [PubMed]
  66. Makhaeva, G.F.; Boltneva, N.P.; Lushchekina, S.V.; Serebryakova, O.G.; Stupina, T.S.; Terentiev, A.A.; Serkov, I.V.; Proshin, A.N.; Bachurin, S.O.; Richardson, R.J. Synthesis, molecular docking and biological evaluation of N,N-disubstituted 2-aminothiazolines as a new class of butyrylcholinesterase and carboxylesterase inhibitors. Bioorg. Med. Chem. 2016, 24, 1050–1062. [Google Scholar] [CrossRef]
  67. Prior, R.L.; Wu, X.; Schaich, K. Standardized Methods for the Determination of Antioxidant Capacity and Phenolics in Foods and Dietary Supplements. J. Agric. Food. Chem. 2005, 53, 4290–4302. [Google Scholar] [CrossRef]
  68. Jiménez, A.; Selga, A.; Torres, J.L.; Julià, L. Reducing Activity of Polyphenols with Stable Radicals of the TTM Series. Electron Transfer versus H-Abstraction Reactions in Flavan-3-ols. Org. Lett. 2004, 6, 4583–4586. [Google Scholar] [CrossRef]
  69. Lespade, L.; Bercion, S. Theoretical investigation of the effect of sugar substitution on the antioxidant properties of flavonoids. Free Radic. Res. 2012, 46, 346–358. [Google Scholar] [CrossRef]
  70. Mohajeri, A.; Asemani, S.S. Theoretical investigation on antioxidant activity of vitamins and phenolic acids for designing a novel antioxidant. J. Mol. Struct. 2009, 930, 15–20. [Google Scholar] [CrossRef]
  71. Amić, D.; Lučić, B. Reliability of bond dissociation enthalpy calculated by the PM6 method and experimental TEAC values in antiradical QSAR of flavonoids. Bioorg. Med. Chem. 2010, 18, 28–35. [Google Scholar] [CrossRef]
  72. Marković, Z.; Milenković, D.; Đorović, J.; Dimitrić Marković, J.M.; Stepanić, V.; Lučić, B.; Amić, D. PM6 and DFT study of free radical scavenging activity of morin. Food Chem. 2012, 134, 1754–1760. [Google Scholar] [CrossRef]
  73. Zhang, D.; Liu, Y.; Chu, L.; Wei, Y.; Wang, D.; Cai, S.; Zhou, F.; Ji, B. Relationship Between the Structures of Flavonoids and Oxygen Radical Absorbance Capacity Values: A Quantum Chemical Analysis. J. Phys. Chem. A 2013, 117, 1784–1794. [Google Scholar] [CrossRef] [PubMed]
  74. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2010.
  75. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  76. Davidson, E.R. Comment on “Comment on Dunning’s correlation-consistent basis sets”. Chem. Phys. Lett. 1996, 260, 514–518. [Google Scholar] [CrossRef]
  77. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  78. Horton, W.; Peerannawar, S.; Török, B.; Török, M. Theoretical and experimental analysis of the antioxidant features of substituted phenol and aniline model compounds. Struct. Chem. 2018, 30, 23–35. [Google Scholar] [CrossRef]
  79. Klein, E.; Lukeš, V.; Cibulková, Z.; Polovková, J. Study of N–H, O–H, and S–H bond dissociation enthalpies and ionization potentials of substituted anilines, phenols, and thiophenols. J. Mol. Struct. 2006, 758, 149–159. [Google Scholar] [CrossRef]
  80. Saqib, M.; Mahmood, A.; Akram, R.; Khalid, B.; Afzal, S.; Kamal, G.M. Density functional theory for exploring the structural characteristics and their effects on the antioxidant properties. J. Pharm. Appl. Chem. 2015, 1, 65–71. [Google Scholar] [CrossRef]
  81. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  82. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  83. Sheldrick, G.M. A short history ofSHELX. Acta Crystallogr. Sec. A 2007, 64, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ishimaru, T. Study on Chemotherapeutics. III. Yakugaku Zasshi 1957, 77, 800–802. [Google Scholar] [CrossRef] [Green Version]
  85. Jo, S.-H.; Ka, E.-H.; Lee, H.-S.; Apostolidis, E.; Jang, H.-D.; Kwon, Y.-I. Comparison of antioxidant potential and rat intestinal α-glucosidases inhibitory activities of quercetin, rutin, and isoquercetin. Int. J. Appl. Res. Nat. Prod. 2009, 2, 52–60. [Google Scholar]
  86. Ellman, G.L.; Courtney, K.D.; Andres, V.; Featherstone, R.M. A new and rapid colorimetric determination of acetylcholinesterase activity. Biochem. Pharmacol. 1961, 7, 88–95. [Google Scholar] [CrossRef]
  87. Sterri, S.H.; Johnsen, B.A.; Fonnum, F. A radiochemical assay method for carboxylesterase, and comparison of enzyme activity towards the substrates methyl [1-14C] butyrate and 4-nitrophenyl butyrate. Biochem. Pharmacol. 1985, 34, 2779–2785. [Google Scholar] [CrossRef]
  88. Pearson, R.G. Hardness of Closed Systems. In Chemical Reactivity Theory: A Density Functional View; Chattaraj, P.K., Ed.; CRC Press: Boca Raton, FL, USA, 2009; pp. 155–162. [Google Scholar]
  89. Cramer, C.J. Essentials of Computational Chemistry: Theories and Models, 2nd ed.; John Wiley & Sons: Chichester, UK, 2004; pp. 330–331. [Google Scholar]
Scheme 1. Reduction of 4-hydroxyiminopyrazolones 1aj.
Scheme 1. Reduction of 4-hydroxyiminopyrazolones 1aj.
Molecules 27 07722 sch001
Figure 1. The structure of compound 2a according to XRD.
Figure 1. The structure of compound 2a according to XRD.
Molecules 27 07722 g001
Figure 2. Antioxidant activity (AOA, ABTS) vs. gap functions for 4-aminopyrazolols 4a–d (a) and (AOA, ABTS) vs. IP functions for 4-aminopyrazolols 4e, 4h–i (b).
Figure 2. Antioxidant activity (AOA, ABTS) vs. gap functions for 4-aminopyrazolols 4a–d (a) and (AOA, ABTS) vs. IP functions for 4-aminopyrazolols 4e, 4h–i (b).
Molecules 27 07722 g002
Scheme 2. The formation of possible radicals in the reactions of 2-hydroxyiminopyrazolones 1a,d,e,j with values of the difference in bond dissociation energies (ΔBDE, kJ/mol), which were evaluated in the approximation of the gas phase (B3LYP/cc-pVDZ method).
Scheme 2. The formation of possible radicals in the reactions of 2-hydroxyiminopyrazolones 1a,d,e,j with values of the difference in bond dissociation energies (ΔBDE, kJ/mol), which were evaluated in the approximation of the gas phase (B3LYP/cc-pVDZ method).
Molecules 27 07722 sch002
Scheme 3. The formation of possible radicals in the reactions of 4-aminopyrazolols 4a,d,e,j * (* conversions are given for 4-aminopyrazolols), with values of the difference in bond dissociation energies (ΔBDE, kJ/mol), which were evaluated in the approximation of the gas phase (B3LYP/cc-pVDZ method).
Scheme 3. The formation of possible radicals in the reactions of 4-aminopyrazolols 4a,d,e,j * (* conversions are given for 4-aminopyrazolols), with values of the difference in bond dissociation energies (ΔBDE, kJ/mol), which were evaluated in the approximation of the gas phase (B3LYP/cc-pVDZ method).
Molecules 27 07722 sch003
Figure 3. Antioxidant activity (AOA, ORAC) vs. BDE functions for 4-hydroxyimino-1-phenyl-pyrazolones 1ej (a) and antioxidant activity (AOA, ORAC) vs. BDE functions for polyfluoroalkyl 4-aminopyrazolols 4ac, 4ei (b).
Figure 3. Antioxidant activity (AOA, ORAC) vs. BDE functions for 4-hydroxyimino-1-phenyl-pyrazolones 1ej (a) and antioxidant activity (AOA, ORAC) vs. BDE functions for polyfluoroalkyl 4-aminopyrazolols 4ac, 4ei (b).
Molecules 27 07722 g003
Table 1. Antioxidant activity of pyrazoles 1 and 4.
Table 1. Antioxidant activity of pyrazoles 1 and 4.
No.CompoundAntioxidant Activity
(Mean ± SEM)
R1R2ABTS
(n = 3; 1 h)
FRAP
(n = 3; 1 h)
ORAC-FL
(n = 3; 2 h)
TEAC *
(IC50, μM) **
TE *TE *
(IC50, μM) **
Molecules 27 07722 i001
1aCF3Ph0.035 ± 0.002
(n.d.)
n.a.n.a.
1bC2F5Ph0.03 ± 0.002
(n.d.)
0.06 ± 0.01n.a.
1cC4F9Ph0.04 ± 0.004
(n.d.)
0.06 ± 0.01n.a.
1dMePh0.16 ± 0.02
(n.d.)
n.a.1.07 ± 0.17
1eCF3H0.23 ± 0.01
(n.d.)
0.07 ± 0.010.98 ± 0.03
1fH(CF2)2H0.15 ± 0.007
(n.d.)
n.a.n.d.
1gC2F5H0.13 ± 0.04
(n.d.)
0.13 ± 0.011.87 ± 0.04
1hC3F7H0.19 ± 0.008
(n.d.)
n.a.2.37 ± 0.09
1iC4F9H0.2 ± 0.01
(n.d.)
n.a.2.41 ± 0.08
1jMeH0.2 ± 0.03
(n.d.)
0.09 ± 0.012.85 ± 0.28
Molecules 27 07722 i002
4aCF3Ph0.75 ± 0.04
(27.2 ± 2.0)
0.89 ± 0.053.49 ± 0.16
(8.8 ± 0.1)
4bC2F5Ph0.6 ± 0.03
(34.1 ± 2.0)
0.75 ± 0.072.12 ± 0.17
4cC4F9Ph0.85 ± 0.04
(23.4 ± 2.1)
0.98 ±0.052.82 ± 0.2
4dMePh0.93 ± 0.03
(23.3 ± 1.9)
0.98 ± 0.084.39 ± 0.04
(6.85 ± 0.7)
4eCF3H0.92 ± 0.04
(21.2 ± 1.8)
1.02 ± 0.020.90 ± 0.19
4fH(CF2)2H0.50 ± 0.02
(33.2 ± 2.3)
0.62 ± 0.01n.d.
4gC2F5H0.73 ± 0.03
(25.7 ± 2.1)
0.88 ± 0.040.54 ± 0.03
4hC3F7H0.78 ± 0.04
(24.8 ± 1.6)
1.03 ± 0.040.63 ± 0.04
4iC4F9H0.45 ± 0.02
(40.8 ± 1.1)
0.75 ± 0.051.83 ± 0.02
4jMeH0.95 ± 0.04
(20.1 ± 1.8)
1.04 ± 0.042.06 ± 0.16
CF3-EDAMolecules 27 07722 i0030.95 ± 0.04
(21.4 ± 0.8)
0.63 ± 0.033.89 ± 0.04
(6.3 ± 0.2)
EDAMolecules 27 07722 i0040.96 ± 0.04
(21.4 ± 1.1)
0.80 ± 0.013.71 ± 0.06
(5.6 ± 0.1)
Trolox1.0
(20.1 ± 1.2)
1.01.0
(23.6 ± 4.4)
Quercetin1.20 ± 0.11
(13.8 ± 0.7)
4.99 ± 0.025.41 ± 0.08
n.d.—not determined; n.a.—not active; * TEAC (Trolox equivalent antioxidant capacity, ABTS) and TE (Trolox equivalent, FRAP, ORAC)—activity of the compounds relative to Trolox (for calculation, see Experimental). ** compound concentration required for 50% reduction of the ABTS radical-cation or peroxyl radical.
Table 2. Esterase profile of pyrazoles 1 and 4.
Table 2. Esterase profile of pyrazoles 1 and 4.
No.CompoundInhibitory Activity against AChE, BChE, and CES
% Inhibition at 20 μM 1 or IC50, μM 2
R1R2AChEBChECES
Molecules 27 07722 i005
1aCF3Phn.a.3n.a.n.a.
1bC2F5Ph11.9 ± 0.9%16.1 ± 1.4%8.2 ± 1.3%
1cC4F9Ph11.5 ± 1.8%5.3 ± 1.0%7.8 ± 1.4%
1dMePh4.9 ± 1.3%6.6 ± 1.3%n.a.
1eCF3Hn.a.n.a.5.9 ± 1.1%
1fH(CF2)2H6.8 ± 1.2%4.7 ± 1.0%3.5 ± 0.9%
1gC2F5Hn.a.n.a.7.8 ± 1.4%
1hC3F7H5.3 ± 1.3%n.a.6.6 ± 1.4%
1iC4F9Hn.a.n.a.n.a.
1jMeHn.a.3.8 ± 1.3%n.a.
Molecules 27 07722 i006
4aCF3Ph7.9 ± 1.5%6.9 ± 1.1%44.9 ± 4.0
4bC2F5Ph4.4 ± 1.1%12.8 ± 1.5%16.6 ± 1.4%
4cC4F9Ph19.2 ± 1.7%10.3 ± 1.2%10.2 ± 0.8
4dMePh6.5 ± 1.2%5.1 ± 1.3%98.7 ± 8.8
4eCF3H9.7 ± 1.5%5.6 ± 1.5%n.a.
4fH(CF2)2H17.3 ± 1.5%n.a.17.5 ± 1.4%
4gC2F5H17.3 ± 1.7%8.9 ± 1.4%6.7 ± 1.2%
4hC3F7Hn.a.3.4 ± 0.9%81.8 ± 7.3
4iC4F9H5.1 ± 0.9%n.a.21.8 ± 1.7
4jMeHn.a.13.9 ± 1.6%4.3 ± 0.9%
CF3-EDAMolecules 27 07722 i007n.a.n.a.7.6 ± 1.2%
EDAMolecules 27 07722 i008n.a.4.6 ± 0.2%8.4 ± 0.7%
1 Percentages correspond to percent inhibition ± SEM at 20 μM, n = 3. 2 Values without units of measurement (numbers printed in bold) correspond to IC50 ± SEM in μM, n = 3. 3 n.a.—not active.
Table 3. Vertical values (IP, EA, η) and BDE(OH) of compounds 1aj, 4aj.
Table 3. Vertical values (IP, EA, η) and BDE(OH) of compounds 1aj, 4aj.
R1CF3C2F5C4F9MeCF3(CF2)2HC2F5C3F7C4F9Me
R2PhPhPhPhHHHHHH
Compounds4-Hydroxyiminopyrazolones
1a1b1c1d1e1f1g1h1i1j
IP, eV8.458.448.447.989.849.619.789.769.779.10
EA, eV1.491.551.600.971.211.111.281.311.340.58
|HOMO-LUMO|
(gap), eV
6.966.896.847.018.638.58.58.458.438.52
η, eV3.483.453.423.504.324.254.254.224.214.26
BDE (OH), kJ/mol315.4314.3314.3314.0314.0320.7313.0312.9313.0312.7
Compounds4-Aminopyrazolols
4a4b4c4d4e4f4g4h4i4j
IP, eV6.816.856.846.417.067.027.097.097.116.56
EA, eV1.331.341.401.050.480.460.590.430.530.12
|HOMO-LUMO|
(gap), eV
−5.47−5.51−5.44−5.36−6.58−6.56−6.50−6.67−6.58−6.43
η, eV2.742.752.722.683.293.283.253.333.293.22
BDE (OH), kJ/mol318.9319.9319.4302.3330.1328.0330.7330.2329.8309.2
EA—electron affinity, IP—ionization potential, HOMO—the highest occupied molecular orbital, LUMO—the lowest unoccupied molecular orbital; η—chemical hardness, BDE—bond dissociation energy.
Table 4. Cytotoxicity of EDA on NHDF.
Table 4. Cytotoxicity of EDA on NHDF.
EDA Inhibitory Capability (IC, %) at Increasing Concentrations
1 µM10 µM100 µM
3.74 ± 0.0519.9 ± 0.832.8 ± 0.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Burgart, Y.V.; Makhaeva, G.F.; Krasnykh, O.P.; Borisevich, S.S.; Agafonova, N.A.; Kovaleva, N.V.; Boltneva, N.P.; Rudakova, E.V.; Shchegolkov, E.V.; Triandafilova, G.A.; et al. Synthesis of 4-Aminopyrazol-5-ols as Edaravone Analogs and Their Antioxidant Activity. Molecules 2022, 27, 7722. https://doi.org/10.3390/molecules27227722

AMA Style

Burgart YV, Makhaeva GF, Krasnykh OP, Borisevich SS, Agafonova NA, Kovaleva NV, Boltneva NP, Rudakova EV, Shchegolkov EV, Triandafilova GA, et al. Synthesis of 4-Aminopyrazol-5-ols as Edaravone Analogs and Their Antioxidant Activity. Molecules. 2022; 27(22):7722. https://doi.org/10.3390/molecules27227722

Chicago/Turabian Style

Burgart, Yanina V., Galina F. Makhaeva, Olga P. Krasnykh, Sophia S. Borisevich, Natalia A. Agafonova, Nadezhda V. Kovaleva, Natalia P. Boltneva, Elena V. Rudakova, Evgeny V. Shchegolkov, Galina A. Triandafilova, and et al. 2022. "Synthesis of 4-Aminopyrazol-5-ols as Edaravone Analogs and Their Antioxidant Activity" Molecules 27, no. 22: 7722. https://doi.org/10.3390/molecules27227722

Article Metrics

Back to TopTop