Detailed Investigation of the Outstanding Peroxyl ... - ACS Publications

Jul 2, 2019 - reaction pathway in lipid and aqueous solutions are reported in. Tables 2 and 3, ... pathways, in principle all of them might contribute...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/jcim

Cite This: J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Detailed Investigation of the Outstanding Peroxyl Radical Scavenging Activity of Two Novel Amino-Pyridinol-Based Compounds Misaela Francisco-Marquez† and Annia Galano*,‡ †

Instituto Politécnico Nacional−UPIICSA, Té 950, Col. Granjas México, C.P. 08400 México City, México Departamento de Química, Universidad Autónoma Metropolitana−Iztapalapa, San Rafael Atlixco 186, Col. Vicentina. Iztapalapa, C.P. 09340, Mexico City, México

Downloaded via UNIV OF SOUTHERN INDIANA on July 22, 2019 at 08:29:48 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: The ability of two novel amino-pyridinol based compounds (NPyr6 and NPyr7) as peroxyl radical scavengers was investigated in silico. The gathered data indicate that they are exceptionally efficient in that role. However, solvent polarity influences their relative efficiency for that purpose. NPyr6 was identified as the best peroxyl radical scavenger in lipid solution, while NPyr7 takes that place in aqueous solution. Both compounds present two acid−base equilibria, which influence their reactivity in aqueous solution. The associated pKa values were estimated. Several reaction mechanisms were explored. Hydrogen transfer from the phenolic group was identified as the chemical route with the highest contribution to the antioxidant behavior of the investigated compounds in both, nonpolar medium and aqueous solution (at 2 ≤ pH ≤ 10). At higher pH other reaction pathways become the most relevant ones. In addition, their bioavailability, cell permeability, safety, and manufacturability were evaluated. According to these, particularly toxicity, NPyr7 seems to be a better candidate for use as an oral drug to fight oxidative stress than NPyr6.



INTRODUCTION Oxidative stress (OS) is a chemical phenomenon with deleterious consequences for human health. It is considered a triggering factor in diverse neurodegenerative disorders,1−16 some types of cancer,17−32 a variety of cardiovascular diseases,33−41 diabetes,15,42,43 and pulmonary44−55 and renal failures.56−64 OS arises when the production of oxidants in biological systems overcomes their consumption.65 Among such oxidants, free radicals are particularly concerning due to their high reactivity and their ability to initiate chain-like reactions, which contribute to propagate molecular damage. Therefore, finding chemicals that act as antioxidants by scavenging free radicals is an active area of research. Such molecules might offer protection by preventing, protecting, or repairing biomolecules.66 Recently, a series of 6-amino-3pyridinols have been investigated as peroxyl (LOO•) radical scavengers and were described as a novel class of phenolic antioxidants.67 In particular, two of them (originally labeled as 7 and 6, and here referred to as NPyr7 and NPyr6, respectively, Scheme 1) were identified as particularly efficient peroxyl radical-trapping chain-breaking antioxidants. According to theoretical calculations, based on O−H bond dissociation energies and ionization potentials, these compounds (Scheme 1) were predicted to be good antioxidants with reasonable air-stability.67 This prediction was experimen© XXXX American Chemical Society

Scheme 1. Structures of NPyr7 and NPyr6 and Site Numbering

tally confirmed. In addition, the inhibition rate constant was measured (based on LOO• trapping and O2 uptake). Such measurements allow classifying NPyr6 and NPyr7 as some of the fastest peroxyl radical-trapping chain-breaking antioxidants reported so far. Later investigations on these compounds also supported this finding. It was reported that NPyr6 and NPyr7 efficiently inhibit styrene autoxidation.68 Moreover, they were found to be 28 and 88 times more efficient for that purpose, respectively, than α-tocopherol (α-TOH). In addition, there Received: June 24, 2019 Published: July 2, 2019 A

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling

for the kinetic calculations. When using this approach is used, the rate constants (k) are calculated using the conventional TST (“transition state theory”),94−96 the 1 M standard state, harmonic vibrational frequencies, and Eckart tunneling.97 The Marcus theory is used to obtain the Gibbs free energy of activation of electron transfer reactions.98 Diffusion is considered, and the calculated rate constants (k) are corrected, accordingly. To that purpose, the Collins−Kimball theory,99 the steady-state Smoluchowski rate constant,100 and the Stokes−Einstein101,102 approach are used.

is evidence that NPyr6 is about 30-fold more reactive toward peroxyl radicals than α-TOH; it protects lipoprotein tryptophan residues in human low-density lipoprotein and does not efficiently mediate lipid peroxidation.69 However, it seems that further investigations on these compounds are still necessary to fully understand their protective actions. To the extent of our knowledge, there is still no full mechanistic study on the role of NPyr6 and NPyr7 as peroxyl radical scavengers. There is no information on the influence of the environment’s polarity on such a role. There is no quantitative data regarding the role of their different acid− base species, which have been shown to be very important, in this context.70 Accordingly, the main goal of the investigation presented here is to address these particular aspects, using a computational approach.



RESULTS AND DISCUSSION Acid Base Equilibria. The first pKa value of NPyr6 and NPyr7 were previously estimated to be higher than 7.0 (7.0− 7.2).68 The values were approximated because the titrations yielded somewhat distorted plots, probably due to decomposition of the anionic species. The fact that these pKa values are higher than that of pyridinium cation (5.25) was expected and attributed to the presence of electron-donating substituents. Here, the PF method was used to estimate two pKa values for each of the investigated molecules. The first one corresponds to the protonated pyridinium moiety and the second one to the phenolic OH. The estimated pKa values, in aqueous solution, at physiological pH (7.4), are reported in Table 1. In the same table, the molar fractions (at the same



COMPUTATIONAL DETAILS Software: the Gaussian 09 package of programs.71 Level of theory: M06-2X/6-311+G(d,p). Solvent effects: SMD (“Solvation Model based on Density”).73 Water and pentyl ethanoate were chosen as solvents to mimic aqueous and lipid solution. The M06-2X choice was made based on previous recommendations, that considering performance for kinetic calculations and reactions of open shell systems.72,74−84 SMD was chosen because its low errors in solvation free energies for solutes of diverse charges.73 The number of imaginary frequencies (if) were used to identify the stationary points as local minima and transition states (if = 0 or if = 1, respectively). IRC (“intrinsic reaction coordinate”) calculations were carried out to confirm that the located transition state (TS) corresponds to the intended reaction pathway. The reported relative energies include thermodynamic corrections at 298.15 K. The pKa values were estimated using the PF (“parameters fitting”) method, which produces only small deviations with respect to experimental measurements.85−87 Once the pKa values are known, the molar fractions of the protonated (H2NPyr+), neutral (HNPyr), and anionic (NPyr−) species of the investigated compounds were calculated, at the pH of interest, as M

f (H NPyr+) = 2

1 + 10

pKa 2

Table 1. Calculated pKa Values of NPyr6 and NPyr7 and Molar Fractions of Their Acid−Base Species at pH = 7.4a pKa1 pKa2 M f(H2NPyr+) × 100 M f(HNPyr) × 100 M f(NPyr−) × 100 a

1 [H ] + 10 pKa 2 + pKa1[H+]2

f (HNPyr) =

10 pKa 2[H+] 1 + 10 pKa 2[H+] + 10 pKa 2 + pKa1[H+]2

M

f (NPyr−) =

10 pKa 2 + pKa1[H+]2 1 + 10 pKa 2[H+] + 10 pKa 2 + pKa1[H+]2

NPyr7

7.8 10.4 71.50 28.47 0.03

6.9 10.5 24.01 75.93 0.06

Expressed as a percent of the total population.

pH) are provided. In addition to the standard procedure associated with the PF and using the recommended parameters for the level of theory used here, a further correction was made. It consisted on estimating the pKa values of phenol and pyridinium cation using the exact methodology and the corresponding deviations with respect to the experimental values. These deviations were in turn been used to improve the calculated pKa values of the molecules of interest. The acid−base species, with largest populations at pH = 7.4 (Table 1), are protonated for NPyr6 (H2NPyr6+) and the product of the first deprotonation for NPyr7 (HNPyr7). However, non-negligible amounts of HNPyr6 and H2NPyr7+ are expected at this pH. The anions of both molecules are predicted to exist only to a minor extent under the same condition. Reaction Mechanisms. Five reaction mechanisms were considered for aqueous solution: • HT: “hydrogen transfer” • RAF: “radical adduct formation” • SET: “single electron transfer” • SPLET: “sequential proton loss electron transfer” • SPLHT: “sequential proton loss hydrogen transfer” • SdPLET: “sequential double-proton loss electron transfer”

+

M

NPyr6

More details on molar fractions calculations can be found elsewhere.88−91 The hydroperoxyl radical (HOO•) was chosen to model the free radical scavenging activity of the target molecules with peroxyl radicals (ROO•). HOO• is the member of the ROO• family with smallest size and has been suggested as crucial to the toxic side effects of aerobic respiration.92 In addition, HOO• can be used also as a model for the reactions of nonhalogenated aliphatic ROO• in kinetic calculations. This strategy was recently validated.93 The QM-ORSA (“quantum mechanism-based test for overall free radical scavenging activity”) protocol88 was used B

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling • SdPLHT: “sequential double-proton loss hydrogen transfer” HT and RAF can take place through different reactions pathways. HT was modeled considering as H donor the O in the phenolic group and any sp3 C atom in NPyr6 and NPyr7. RAF was modeled considering any sp2 C atom in the investigated molecules as reaction sites. Regarding SET, SPLET, and SdPLET, they differ in the particular acid−base species that is acting as the electron donor (H2NPyr+, HNPyr, and NPyr−, respectively). The same applies for HT, SPLHT, and SdPLHT, which correspond to hydrogen transfers from the different acid−base species. For reactions in lipid media those mechanisms involving charged species were not included because such species are expected to exist to a significant extent only when the medium is a polar and protic solvent. Thus, in lipid solution only two reaction mechanism were considered as possible: HT and RAF, and the molecules of interest were modeled as neutral species (HNPyr). Thermochemistry. The Gibbs free energies (ΔG) of each reaction pathway in lipid and aqueous solutions are reported in Tables 2 and 3, respectively. Only two of them were found to

exergonicity in both cases (Table 2). Neutral species (HNPyr), in aqueous solution, have a similar behavior (Table 3). However, the exergonicity increases to some extent with the polarity of the solvent. For the charged species the number of thermochemically viable reaction pathways increases to 3 for H2NPyr6+, to 4 for H2NPyr7+, and to 5 for NPyr6− and NPyr7− (Table 3). For the protonated species, which have the phenolic moiety, HT from this site is the most exergonic channel. On the other hand, in the anions yielded by deprotonation of the phenolic group, this pathway is not possible. The most negative ΔG value for the reactions of NPyr6− and NPyr7− with HOO• corresponds to the SdPLET mechanism. Since the reactions of HOO• with all the acid−base species of NPyr6 and NPyr7 involves some thermochemically viable pathways, in principle all of them might contribute to the peroxyl radical scavenging activity of the investigated aminopyridinols. However, kinetic calculations are required to estimate the extent of such contributions. Kinetics. The reaction pathways identified as endergonic in the previous section were excluded from the kinetic calculations. It is not expected that the products yielded through these pathways will be experimentally observed, even if they might be formed at a significant rate. On the other hand, moderately endergonic pathways might be non-negligible, provided that their products evolve fast enough into other species through reactions that are significantly exergonic. These may be the case for the SPLET mechanism, which yields radical anions that are expected to be highly reactive. Thus, this mechanism was taken into consideration in the kinetic analyses, despite of being moderately endergonic. The most relevant structural features of the TS are shown in Figures 1−4. Their imaginary frequencies are reported in Table S1 (Supporting Information), while the corresponding tunneling corrections are provided in Table S2. The energy barriers of all the reactions pathways considered for kinetics are reported in Table S3. Total rate constants in lipid solution, and for each acid−base species in aqueous solution, were calculated as

Table 2. Gibbs Free Energies of Reaction (ΔG, in kcal/mol, at 298.15 K) for the Modeled Pathways, in Lipid Solution HT, OH site HT, 2a site HT, 4a site HT, 7a site HT, 8 site HT, 9 site HT, 10 site RAF, C2 site RAF, C3 site RAF, C4 site RAF, C5 site RAF, C6 site

HNPyr6

HNPyr7

−14.03 3.56 6.11 3.93 2.48 11.30 −2.00 9.39 9.99 19.05 11.16 16.88

−14.66 3.39 7.32 4.94 0.77 −3.53 8.10 10.66 20.55 10.97 20.60

be thermochemically viable in lipid solution for both NPyr6 (HT from the OH site and HT from site 10) and NPyr7 (HT from the OH site and HT from site 9). The reaction pathway involving HT from the phenolic site has the largest

HNPyr k total,PE =

n1

m1

HNPyr HNPyr + ∑ kRAF ∑ kHT i

i=1

j

j=1

Table 3. Gibbs Free Energies of Reaction (ΔG, in kcal/mol, at 298.15 K) for the Modeled Pathways, in Aqueous Solution

HT, OH site HT, 2a site HT, 4a site HT, 7a site HT, 8 site HT, 9 site HT, 10 site RAF, C2 site RAF, C3 site RAF, C4 site RAF, C5 site RAF, C6 site SET SPLET SdPLET

H2NPyr6+

HNPyr6

NPyr6−

−9.42 −1.53 3.38 5.79 2.55 9.83 −2.65 0.93 11.59 17.16 14.76 16.49 25.68

−17.11 1.34 3.73 1.88 0.30 8.99 −4.04 7.16 9.24 18.53 8.90 12.20

−3.88 −2.12 2.24 1.66 9.41 −4.06 −1.77 9.56 4.42 15.38 5.99

9.95

H2NPyr7+

HNPyr7

NPyr7−

−11.57 −2.68 4.33 4.18 3.98 −4.68

−17.68 0.76 3.94 3.35 1.23 −3.93

−5.36 −0.75 3.75 2.43 −3.57

−1.56 11.12 16.14 8.44 13.01 23.61

6.40 8.46 19.44 7.10 14.99

−4.49 9.04 4.53 10.29 4.07

6.73 −13.58 C

−15.16 DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling

Figure 1. Optimized geometries of the transition states of neutral species, in lipid solution. Distances are reported in angstroms. H2 NPyr + k total,W

n2

=



H 2NPyr + kHT + i

i=1

HNPyr k total,W

n1

=







H 2NPyr + kRAF j

Figure 3. Optimized geometries of the transition states of the neutral species (HNPyr), in aqueous solution. Distances are reported in angstroms.

+ k SET

j=1

HNPyr kHT + i

i=1

NPyr k total,W =

m2

m1



HNPyr kRAF j

The difference between these two expressions is a consequence of the acid−base equilibrium of HOO• in aqueous solution (pKa = 4.8).92 In lipid solution (Table 4) NPyr6 reacts faster than NPyr7 with peroxyl radicals. In this medium, NPyr6 was identified as the best peroxyl radical scavenger identified so far (Table S4, Supporting Information). This statement is based on comparisons of total rate constants, calculated using similar computational protocols.66,103,104 NPyr7 is also excellent for this purpose and surpassed only by lycopene and torulene, which react 1.9 and 1.1 times faster with HOO• in nonpolar media (koverall = 1.69 × 106 and 9.47 × 105 M−1 s−1, respectively).105 Under the same conditions, NPyr6 and NPyr7 react 700 and 278 times faster than Trolox.106 They are also faster in scavenging peroxyl radicals than gallic acid,107 resveratrol,108 and ascorbic acid.109 HT is the only mechanism contributing to the peroxyl radical scavenging activity of NPyr6 and NPyr7, in lipid solution, with the phenolic site accounting for ∼100% of such

+ k SPLET

j=1

n3

m3



NPyr NPyr + ∑ kRAF ∑ kHT i



j

i=1

+ k SdPLET

j=1

Overall rate constants, in aqueous solution at each pH of interest, for the reactions with model ROO• and HOO• were calculated as +

pH

pH

HNPyr H 2NPyr W,pH M koverall,ROO f (H NPyr+) k total,W + M f (HNPyr) k total,W • = 2

pH



NPyr + M f (NPyr−) k total,W

pH

W,pH W,pH M koverall,HOO f (HOO•) koverall,ROO • = •

Figure 2. Optimized geometries of the transition states of the protonated species (H2NPyr+), in aqueous solution. Distances are reported in angstroms. D

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling

Figure 4. Optimized geometries of the transition states of the anionic species (NPyr6−), in aqueous solution. Distances are reported in angstroms.

The reactions of both molecules with HOO• are predicted to be slower than with other ROO• (Table 5), at pH= 7.4, which is a consequence of the small fraction of HOO• that is present at this pH (∼0.25%). NPyr6 and NPyr7 are both excellent for scavenging peroxyl radicals, in aqueous solution at physiological pH. In fact, NPyr7 is the best antioxidant identified so far for that purpose (Table S6, Supporting Information), while NPyr6 is only surpassed by 3,5-dihydroxy4-methoxybenzyl alcohol (DHMBA) and piceatannol (koverall,ROO• = 1.40 × 109 and 1.13 × 109 M−1 s−1, respectively).110,111 The finding that NPyr6 and NPyr7 are outstanding peroxyl radical scavengers, regardless of the solvent polarity, might be attributed to the presence of a nitrogen atom in the aromatic ring. It is expected to increase the electronic density, consequently, to stabilize the radical products yielded through the reactions with free radicals. In addition, phenolic compounds with increased electrodonating capabilities are expected to be more reactive toward oxidants including Ocentered free radicals, such as ROO•. The acid−base equilibria of the investigated aminopyridinols have a significant influence on the kinetics of their reactions with peroxyl radicals and, consequently, on their capability as peroxyl radical scavengers. This activity increases with the deprotonation degree (Table 5), H2NPyr+ being significantly less efficient for that purpose than HNPyr and

Table 4. Rate Constants (k, M−1 s−1) of the Reactions between Amino-pyridinols and Peroxyl Radicals, in Lipid Solution kHNPyr6 total,PE HT, OH site HT, 9 site HT, 10 site total

kHNPyr7 total,PE

2.38 × 10

8.78 × 105 7.16 × 100

6

3.11 × 101 2.38 × 106

8.78 × 105

activity (Table S5, Supporting Information). Thus, at least, in this medium the phenolic moiety is identified as responsible for the antioxidant behavior of the investigated compounds. In aqueous solution, the chemistry involved in the peroxyl radical scavenging activity of NPyr6 and NPyr7 becomes significantly more complex, mainly as a consequence of the acid−base behavior of the chemicals involved in the process. Branching ratios for the different reaction pathways are reported in Table S5 (Supporting Information). According to the overall calculated rate constants (Table 5), both NPyr6 and NPyr7 are predicted to react with ROO• at (or near to) diffusion-limited rates, in aqueous solution at physiological pH. Under such conditions, the reaction involving NPyr7 is predicted to be about 2.7 times faster than that involving NPyr6. Both results are in agreement with previous experimental evidence.67

Table 5. Rate Constants (k, M−1 s−1) of the Reactions between Amino-Pyridinol Species and Peroxyl Radicals, in Aqueous Solution

HT, OH site HT, 2a site HT, 4a site HT, 9 site HT, 10 site RAF, C2 site SET SPLET SdPLET total overall, model ROO•, pH = 7.4 overall, HOO•, pH = 7.4

H2NPyr6+

HNPyr6

2.28 × 10 1.62 × 10−1

2.70 × 10

2.83 × 10

9.65 × 10

4

0

NPyr6− 9

HNPyr7

1.05 × 10 2.13 × 10−1

2.70 × 10

7.56 × 100

1.46 × 102

5

2.27 × 109 1.75 × 107 8.19 × 10 2.13 × 109

1

NPyr7−

H2NPyr7+

9

7

1.79 × 10−7

9.73 × 102 2.65 × 10−5

3.13 × 105

2.74 × 109 4.27 × 105

7.42 × 10 1.19 × 1010 9

2.28 × 104

2.65 × 109 5.13 × 106 6.17 × 108

2.70 × 109 7.71 × 108 1.93 × 106 E

1.06 × 105

2.70 × 109 2.06 × 109 5.16 × 106

7.37 × 109 1.34 × 1010

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling NPyr−. The neutral and anionic fractions both react at diffusion limited rates with peroxyl radicals. The values of the total rate constants for each acid−base species and the overall rate constants at pH values ranging from 2.0 to 12.0 are provided for the reactions of NPyr6 and NPyr7 with model ROO• and HOO• in Tables S8 and S9 (Supporting Information), respectively. To facilitate discussion, the dependence of koverall with pH is also presented in Figure 5. The contributions (%) of acid−base species to the

at different pHs, are provided in Tables S11 and S12 (Supporting Information) for NPyr6 and NPyr7, respectively. It was found that NPyr7 reacts faster with peroxyl radicals than NPyr6, for most of the investigated pH range (Figure 5). The exception occurs from pH ∼ 8.5 to pH ∼ 10.4, where both compounds react almost equally fast with these radicals. For the reactions with HOO•, the overall rate constant for NPyr7 surpasses the 107 M−1 s−1 value in the 4.8−7.0 pH range. The maximum reactivity of NPyr6 is reached in the same pH interval, although the values of the corresponding overall rate constants are 6 to 8 times lower. For the reactions with HOO•, on the other hand, the rate constants of both NPyr7 and NPyr6 systematically increases with the pH. The value of koverall is larger than 107 M−1 s−1 at pH values higher than ∼4.4 and ∼5.4, for NPyr7 and NPyr6, respectively. The contributions of the different acid−base species to the overall reactivity of the investigated amino-pyridinols toward peroxyl radicals (Table S10, Supporting Information) are not exactly in line with their relative abundances (Figure 6). The neutral species (HNPyr) are the most abundant ones at pH values ranging from 7.8 to 10.4 and from 6.9 to 10.5 for NPyr6 and Npyr7, respectively. In spite of this, HNPyr species are the key players in the peroxyl radical scavenging activity exhibited by both compounds in a wider range (3.0 ≤ pH ≤ 9.5). In other words, although at pH = 6 (for example) the protonated species (H2Pyr+) are the most abundant ones, the antioxidant activity is ruled by HNPyr for both NPyr6 and Npyr7. In addition, there are several reaction pathways for each acid−base species of the investigated amino-pyridinols that might contribute to their antioxidant behavior as peroxyl radical scavengers. The branching ratios corresponding to

Figure 5. Dependence of kinetics with pH for the reactions between amino-pyridinols and peroxyl radicals, in aqueous solution.

overall rate constants, at different pHs, are reported in Table S10 (Supporting Information). The branching ratios (%) of each reaction channel, with respect to the overall rate constant,

Figure 6. Influence of pH on molar fractions and contributions of acid−base species to the overall rate constants. F

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling

Figure 7. Dependence with pH of branching ratios (%, vs koverall) of the significant reaction pathways for the reactions between amino-pyridinols and peroxyl radicals, in aqueous solution: (A) NPyr6, (B) NPyr7.

Bioavailability and cell permeability may be affected by several physicochemical properties. To explore the behavior of NPyr7 and Npyr6, in this context, several parameters were estimated and used as indicators for ADME (“absorption, distribution, metabolism, and excretion”) properties. They allow checking if the target molecules satisfy Lipinski’s,112 Ghose’s,113 and Veber’s114 criteria, according to the target values shown in Table 6. The ADME properties logP, XAt, MW, HB A , HB D , and RB were obtained using the Molinspiration Property Calculation Service, while MR and PSA were estimated with the DruLiTo software. The values of the calculated ADME properties and toxicity descriptors are reported in Table 7. Five commonly used medical drugs were also included in this table for comparison purposes. It was found that most of the ADME properties for NPyr6 and NPyr7 are within the target values. The exceptions are XAt for both of them (XAtNPyr6 = 14 and XAtNPyr7 = 13) and M R for NPyr6 (MRNPyr6 = 38.18). Regarding XAt, it should be considered that most of the reference medical drugs also have

these pathways, in the 2−12 pH range were estimated (Tables S11 and S12, Supporting Information). Those contributing to a non-negligible extension to the overall activity of NPyr6 and NPyr7 are plotted in Figure 7. It was found that for neutral species (HNPyr) the only relevant pathway is the HT from the phenolic OH, along the whole pH range investigated here. The same applies for the protonated species (H2NPyr+). On the contrary, there are several reaction pathways that significantly contribute to the ROO• scavenging activity of the anionic species (NPyr−), albeit the contributions of the SdPLET route are the largest, by far, for this fraction. Other Relevant Considerations. According to the data discussed in the previous section, both NPyr7 and Npyr6 are highly promising antioxidants. Moreover, their efficiency as peroxyl radical scavengers seems to be among the highest reported so far. Thus, they might be used to protect biomolecules from these oxidants and perhaps as oral antioxidants. However, to even consider such a possibility there are other important factors to take into account. G

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling

suggest that the synthesis of these compounds should be of medium complexity (or medium-easy, since they are close to the first threshold). Therefore, this is not expected to be an issue. Moreover, the SA value of omeprazole (which is commercially viable) is higher (4.15). Probably the most critical aspect regarding potential oral drugs is toxicity. NPyr6 was found to have positive mutagenicity (M = 0.53), while NPyr7 is predicted to be nonmutagenic (M = 0.18). Although there are currently used drugs, such as diclofenac with similar M values (also 0.53), it is preferable to avoid this risk. The LD50 values also indicate that the toxicity of NPyr7 is lower than that of NPyr6. In addition, the LD50 for NPyr7 is higher than the average of the currently used medical drugs, used here as references, while that of NPyr6 is lower. Therefore, according to the estimated toxicity descriptors, NPyr7 seems to have higher potential than NPyr6 as an oral drug to be used for fighting OS.

Table 6. ADME Properties and Their Target Values (TVal), Chosen to Satisfy Lipinski’s, Ghose’s, and Veber’s Criteria property

acronym

TVal

partition coefficient octanol/water polar surface area number of non-hydrogen atoms molecular weight number of H bond acceptors number of H bond donors number of rotatable bonds molar refractivity

logP PSA X At MW HBA HBD RB M R

−0.4−5.0 ≤140 20−70 160−480 ≤10 ≤5 ≤10 40−130

units Å2 g/mol

m3/mol

values lower than 20, omeprazole being the only exception. Since for NPyr6 XAt is the only violation to Lipinski’s, Ghose’s, and Veber’s criteria, while there are two violations for NPyr7, NPyr6 seems to be a better candidate to be used as an oral drug to fight OS. However, all the above-mentioned criteria are general guidelines, not strict rules. For example, paracetamol present two violations of these criteria (XAt and MW) and is a widely used oral drug. In addition, viable drugs must also have other important features, including safety and manufacturability.115 Thus, these aspects were also investigated here. Two toxicity descriptors were used in this work, namely: • M: known as Ames mutagenicity. A chemical is positive if it induces revertant colony growth in any strain of Salmonella typhimurium. When M < 0.5, the chemical is considered not mutagenic. • LD50: amount of chemical per body weight (mg/kg) leading to 50% of rats dying. The larger the LD50 value, the lower the toxicity. They were estimated using T.E.S.T. (Toxicity Estimation Software Tool), which makes predictions based on QSAR (“quantitative structure activity relationships”), and computed with the consensus method.116 Synthetic accessibility (SA) was used to assess manufacturability. It was estimated with the SYLVIA-XT program,117 which uses several contributing criteria, scaled and weighted to produce SA values between 1 and 10, with a larger value indicating a compound more difficult to synthesize. Here, the following classification is used: easy (SA ≤ 3), medium (3 < SA ≤ 6), and hard (SA > 6). Regarding manufacturability, both NPyr6 and NPyr7 have similar SA values (3.58 and 3.51, respectively). These values



CONCLUSIONS The peroxyl radical scavenging activity of two novel aminopyridinol based compounds was investigated here (namely, NPyr6 and NPyr7). They were found to be exceptionally good for that purpose. NPyr6 is at the top of the peroxyl radical scavengers, in lipid solution, identified so far; while NPyr7 was identified as the most efficient in aqueous solution. These compounds present two acid−base equilibria, which influence their reactivity in aqueous solution. The associated pKa values were estimated and are reported here for the first time. Different reaction mechanisms were explored, and H transfer from the phenolic group was identified as the chemical pathway with the highest contribution to the antioxidant behavior of the investigated compounds. This was the case for nonpolar medium and for most of the pH range investigated here for aqueous solution. Only at rather high pH other reaction pathways become the most relevant ones. According to the thermochemical and kinetic data gathered here, both NPyr7 and Npyr6 are both highly promising antioxidants that might be used as oral antioxidants to protect biomolecules for OS. To evaluate the likeliness of such a possibility other important factors were considered (bioavailability, cell permeability, safety, and manufacturability). For that purpose several ADME properties, toxicity descriptors. and synthetic availability were estimated. According to them,

Table 7. Estimated ADME Properties and Toxicity Descriptors for the Investigated Amino-pyridinol Thiols and Some Currently Used Oral Medical Drugs HNPyr6

HNPyr7

melatonin

logP PSA X At MW HBA HBD RB M R

2.18 36.36 14 192.26 3 1 0 38.18

1.66 23.47 13 178.24 3 1 0 45.24

1.45 54.12 17 232.28 4 2 4 66.28

M LD50

0.53 665.83

0.18 1020.63

0.05 1298.11

SA

3.58

3.51

2.46

aspirin ADME 1.43 63.6 13 180.16 4 1 3 43.95 Toxicity 0.43 757.21 Manufacturability 2.20 H

paracetamol (acetaminophen)

diclofenac

omeprazole

0.68 49.33 11 151.16 2 0 1 40.83

4.57 49.33 19 296.15 3 2 4 75.46

2.41 77.11 24 345.42 6 1 5 93.81

0.43 1806.98

0.53 244.02

0.40 1037.14

2.00

2.68

4.15

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling

(12) Yan, M. H.; Wang, X.; Zhu, X. Mitochondrial Defects and Oxidative Stress in Alzheimer Disease and Parkinson Disease. Free Radical Biol. Med. 2013, 62, 90−101. (13) Cheignon, C.; Tomas, M.; Bonnefont-Rousselot, D.; Faller, P.; Hureau, C.; Collin, F. Oxidative Stress and the Amyloid Beta Peptide in Alzheimer’s Disease. Redox Biol. 2018, 14, 450−464. (14) Pu, Z.; Xu, W.; Lin, Y.; He, J.; Huang, M. Oxidative Stress Markers and Metal Ions Are Correlated with Cognitive Function in Alzheimer’s Disease. Am. J. Alzheimers Dis Other Demen 2017, 32, 353−359. (15) Umeno, A.; Biju, V.; Yoshida, Y. In Vivo Ros Production and Use of Oxidative Stress-Derived Biomarkers to Detect the Onset of Diseases Such as Alzheimer’s Disease, Parkinson’s Disease, and Diabetes. Free Radical Res. 2017, 51, 413−427. (16) Jiang, T.; Sun, Q.; Chen, S. Oxidative Stress: A Major Pathogenesis and Potential Therapeutic Target of Antioxidative Agents in Parkinson’s Disease and Alzheimer’s Disease. Prog. Neurobiol. 2016, 147, 1−19. (17) Wang, J.; Li, J. Z.; Lu, A. X.; Zhang, K. F.; Li, B. J. Anticancer Effect of Salidroside on A549 Lung Cancer Cells through Inhibition of Oxidative Stress and Phospho-P38 Expression. Oncol. Lett. 2014, 7, 1159−1164. (18) Fan, C.; Pan, Y.; Yang, Y.; Di, S.; Jiang, S.; Ma, Z.; Li, T.; Zhang, Z.; Li, W.; Li, X.; Reiter, R. J.; Yan, X. Hdac1 Inhibition by Melatonin Leads to Suppression of Lung Adenocarcinoma Cells Via Induction of Oxidative Stress and Activation of Apoptotic Pathways. J. Pineal Res. 2015, 59, 321−333. (19) Tekiner-Gulbas, B.; Westwell, A. D.; Suzen, S. Oxidative Stress in Carcinogenesis: New Synthetic Compounds with Dual Effects Upon Free Radicals and Cancer. Curr. Med. Chem. 2013, 20, 4451− 4459. (20) Valko, M.; Izakovic, M.; Mazur, M.; Rhodes, C. J.; Telser, J. Role of Oxygen Radicals in DNA Damage and Cancer Incidence. Mol. Cell. Biochem. 2004, 266, 37−56. (21) Valko, M.; Rhodes, C. J.; Moncol, J.; Izakovic, M.; Mazur, M. Free Radicals, Metals and Antioxidants in Oxidative Stress-Induced Cancer. Chem.-Biol. Interact. 2006, 160, 1−40. (22) Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M. T. D.; Mazur, M.; Telser, J. Free Radicals and Antioxidants in Normal Physiological Functions and Human Disease. Int. J. Biochem. Cell Biol. 2007, 39, 44−84. (23) Halliwell, B. Biochemistry of Oxidative Stress. Biochem. Soc. Trans. 2007, 35, 1147−1150. (24) Boyd, N. F.; McGuire, V. The Possible Role of Lipid Peroxidation in Breast Cancer Risk. Free Radical Biol. Med. 1991, 10, 185−190. (25) Nelson, R. L. Dietary Iron and Colorectal Cancer Risk. Free Radical Biol. Med. 1992, 12, 161−168. (26) Knekt, P.; Reunanen, A.; Takkunen, H.; Aromaa, A.; Heliovaara, M.; Hakulinen, T. Body Iron Stores and Risk of Cancer. Int. J. Cancer 1994, 56, 379−382. (27) Omenn, G. S.; Goodman, G. E.; Thornquist, M. D.; Balmes, J.; Cullen, M. R.; Glass, A.; Keogh, J. P.; Meyskens Jr, F. L.; Valanis, B.; Williams Jr, J. H.; Barnhart, S.; Hammar, S. Effects of a Combination of Beta Carotene and Vitamin a on Lung Cancer and Cardiovascular Disease. N. Engl. J. Med. 1996, 334, 1150−1155. (28) Willcox, J. K.; Ash, S. L.; Catignani, G. L. Antioxidants and Prevention of Chronic Disease. Crit. Rev. Food Sci. Nutr. 2004, 44, 275−295. (29) Granados-Principal, S.; El-Azem, N.; Pamplona, R.; RamirezTortosa, C.; Pulido-Moran, M.; Vera-Ramirez, L.; Quiles, J. L.; Sanchez-Rovira, P.; Naudí, A.; Portero-Otin, M.; Perez-Lopez, P.; Ramirez-Tortosa, M. Hydroxytyrosol Ameliorates Oxidative Stress and Mitochondrial Dysfunction in Doxorubicin-Induced Cardiotoxicity in Rats with Breast Cancer. Biochem. Pharmacol. 2014, 90, 25− 33. (30) Nichols, H. B.; Anderson, C.; White, A. J.; Milne, G. L.; Sandler, D. P. Oxidative Stress and Breast Cancer Risk in Premenopausal Women. Epidemiology 2017, 28, 667−674.

particularly toxicity, NPyr7 seems to be a better candidate for being used as an oral drug to fight OS than NPyr6.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jcim.9b00517. Imaginary frequencies of the HT and RAF transition states. Tunneling corrections for the HT reaction pathways. Gibbs energy of activation for the different reaction pathways. Kinetic data for the reactions of other antioxidants with peroxyl radicals. Branching ratios. Dependence with pH of the rate constants. Contributions of acid−base species to the overall rate constant at different pHs (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected], [email protected]. Phone: (52) 5558044600. ORCID

Annia Galano: 0000-0002-1470-3060 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We acknowledge the “Laboratorio de Visualización y Cómputo Paralelo” at Universidad Autónoma Metropolitana−Iztapalapa. This investigation is inserted in research projects IFC-2016/ 1828 and SIP 20182008.



REFERENCES

(1) Halliwell, B. Role of Free Radicals in the Neurodegenerative Diseases: Therapeutic Implications for Antioxidant Treatment. Drugs Aging 2001, 18, 685−716. (2) López, N.; Tormo, C.; De Blas, I.; Llinares, I.; Alom, J. Oxidative Stress in Alzheimer’s Disease and Mild Cognitive Impairment with High Sensitivity and Specificity. J. Alzheimer's Dis. 2013, 33, 823−829. (3) Pohanka, M. Alzheimer’s Disease and Oxidative Stress: A Review. Curr. Med. Chem. 2013, 21, 356−364. (4) Pimentel, C.; Batista-Nascimento, L.; Rodrigues-Pousada, C.; Menezes, R. A. Oxidative Stress in Alzheimer’s and Parkinson’s Diseases: Insights from the Yeast Saccharomyces Cerevisiae. Oxid. Med. Cell. Longevity 2012, 2012, 132146. (5) Sayre, L. M.; Smith, M. A.; Perry, G. Chemistry and Biochemistry of Oxidative Stress in Neurodegenerative Disease. Curr. Med. Chem. 2001, 8, 721−738. (6) Chen, Z.; Zhong, C. Oxidative Stress in Alzheimer’s Disease. Neurosci. Bull. 2014, 30, 271−281. (7) Meraz-Ríos, M. A.; Franco-Bocanegra, D.; Toral Rios, D.; Campos-Peña, V. Early Onset Alzheimer’s Disease and Oxidative Stress. Oxid. Med. Cell. Longevity 2014, 2014, 375968. (8) Rosini, M.; Simoni, E.; Milelli, A.; Minarini, A.; Melchiorre, C. Oxidative Stress in Alzheimer’s Disease: Are We Connecting the Dots? J. Med. Chem. 2014, 57, 2821−2831. (9) Schrag, M.; Mueller, C.; Zabel, M.; Crofton, A.; Kirsch, W. M.; Ghribi, O.; Squitti, R.; Perry, G. Oxidative Stress in Blood in Alzheimer’s Disease and Mild Cognitive Impairment: A MetaAnalysis. Neurobiol. Dis. 2013, 59, 100−110. (10) Zhao, Y.; Zhao, B. Oxidative Stress and the Pathogenesis of Alzheimer’s Disease. Oxid. Med. Cell. Longevity 2013, 2013, 316523. (11) Eskici, G.; Axelsen, P. H. Copper and Oxidative Stress in the Pathogenesis of Alzheimer’s Disease. Biochemistry 2012, 51, 6289− 6311. I

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling (31) Mikuła-Pietrasik, J.; Uruski, P.; Pakuła, M.; Maksin, K.; Szubert, S.; Woźniak, A.; Naumowicz, E.; Szpurek, D.; Tykarski, A.; Ksia̧zė k, K. Oxidative Stress Contributes to Hepatocyte Growth Factor-Dependent Pro-Senescence Activity of Ovarian Cancer Cells. Free Radical Biol. Med. 2017, 110, 270−279. (32) Saha, S. K.; Lee, S. B.; Won, J.; Choi, H. Y.; Kim, K.; Yang, G. M.; Dayem, A. A.; Cho, S. G. Correlation between Oxidative Stress, Nutrition, and Cancer Initiation. Int. J. Mol. Sci. 2017, 18, 1544. (33) Matsuda, M.; Shimomura, I. Roles of Adiponectin and Oxidative Stress in Obesity-Associated Metabolic and Cardiovascular Diseases. Rev. Endocr. Metab. Disord. 2014, 15, 1−10. (34) Eren, E.; Ellidag, H. Y.; Cekin, Y.; Ayoglu, R. U.; Sekercioglu, A. O.; Yilmaz, N. Heart Valve Disease: The Role of Calcidiol Deficiency, Elevated Parathyroid Hormone Levels and Oxidative Stress in Mitral and Aortic Valve Insufficiency. Redox Rep. 2014, 19, 34−39. (35) Zampetaki, A.; Dudek, K.; Mayr, M. Oxidative Stress in Atherosclerosis: The Role of Micrornas in Arterial Remodeling. Free Radical Biol. Med. 2013, 64, 69−77. (36) Steinberg, D. Antioxidants and Atherosclerosis. A Current Assessment. Circulation 1991, 84, 1420−1425. (37) Panasenko, O. M.; Vol’Nova, T. V.; Azizova, O. A.; Vladimirov, Y. A. Free Radical Modification of Lipoproteins and Cholesterol Accumulation in Cells Upon Atherosclerosis. Free Radical Biol. Med. 1991, 10, 137−148. (38) Al-Aubaidy, H. A.; Jelinek, H. F. Oxidative Stress and Triglycerides as Predictors of Subclinical Atherosclerosis in Prediabetes. Redox Rep. 2014, 19, 87−91. (39) Aljunaidy, M. M.; Morton, J. S.; Cooke, C. L. M.; Davidge, S. T. Prenatal Hypoxia and Placental Oxidative Stress: Linkages to Developmental Origins of Cardiovascular Disease. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2017, 313, R395−R399. (40) Kelly, F. J.; Fussell, J. C. Role of Oxidative Stress in Cardiovascular Disease Outcomes Following Exposure to Ambient Air Pollution. Free Radical Biol. Med. 2017, 110, 345−367. (41) Nair, N.; Gongora, E. Oxidative Stress and Cardiovascular Aging: Interaction between Nrf-2 and Adma. Curr. Cardiol. Rev. 2017, 13, 183−188. (42) Gerber, P. A.; Rutter, G. A. The Role of Oxidative Stress and Hypoxia in Pancreatic Beta-Cell Dysfunction in Diabetes Mellitus. Antioxid. Redox Signaling 2017, 26, 501−518. (43) Chou, S. T.; Tseng, S. T. Oxidative Stress Markers in Type 2 Diabetes Patients with Diabetic Nephropathy. Clin. Exp. Nephrol. 2017, 21, 283−292. (44) Zhao, W.; Gan, X.; Su, G.; Wanling, G.; Li, S.; Hei, Z.; Yang, C.; Wang, H. The Interaction between Oxidative Stress and Mast Cell Activation Plays a Role in Acute Lung Injuries Induced by Intestinal Ischemia-Reperfusion. J. Surg. Res. 2014, 187, 542−552. (45) MacNee, W. Oxidative Stress and Lung Inflammation in Airways Disease. Eur. J. Pharmacol. 2001, 429, 195−207. (46) Caramori, G.; Papi, A. Oxidants and Asthma. Thorax 2004, 59, 170−173. (47) Guo, R. F.; Ward, P. A. Role of Oxidants in Lung Injury During Sepsis. Antioxid. Redox Signaling 2007, 9, 1991−2002. (48) Hoshino, Y.; Mishima, M. Redox-Based Therapeutics for Lung Diseases. Antioxid. Redox Signaling 2008, 10, 701−704. (49) Shadab, M.; Agrawal, D. K.; Aslam, M.; Islam, N.; Ahmad, Z. Occupational Health Hazards among Sewage Workers: Oxidative Stress and Deranged Lung Functions. J. Clin. Diagnostic Res. 2014, 8, BC11−BC13. (50) Ghasemzadeh, N.; Patel, R. S.; Eapen, D. J.; Veledar, E.; Kassem, H. A.; Manocha, P.; Khayata, M.; Zafari, A. M.; Sperling, L.; Jones, D. P.; Quyyumi, A. A. Oxidative Stress Is Associated with Increased Pulmonary Artery Systolic Pressure in Humans. Hypertension 2014, 63, 1270−1275. (51) Cheresh, P.; Kim, S. J.; Tulasiram, S.; Kamp, D. W. Oxidative Stress and Pulmonary Fibrosis. Biochim. Biophys. Acta, Mol. Basis Dis. 2013, 1832, 1028−1040.

(52) Pandey, R.; Singh, M.; Singhal, U.; Gupta, K. B.; Aggarwal, S. K. Oxidative/Nitrosative Stress and the Pathobiology of Chronic Obstructive Pulmonary Disease. J. Clin. Diagn. Res. 2013, 7, 580−588. (53) Lanzetti, M.; Da Costa, C. A.; Nesi, R. T.; Barroso, M. V.; Martins, V.; Victoni, T.; Lagente, V.; Pires, K. M. P.; E Silva, P. M. R.; Resende, A. C.; Porto, L. C.; Benjamim, C. F.; Valença, S. S. Oxidative Stress and Nitrosative Stress Are Involved in Different Stages of Proteolytic Pulmonary Emphysema. Free Radical Biol. Med. 2012, 53, 1993−2001. (54) Zuo, L.; Hallman, A. H.; Yousif, M. K.; Chien, M. T. Oxidative Stress, Respiratory Muscle Dysfunction, and Potential Therapeutics in Chronic Obstructive Pulmonary Disease. Front. Biol. 2012, 7, 506− 513. (55) Kratzer, E.; Tian, Y.; Sarich, N.; Wu, T.; Meliton, A.; Leff, A.; Birukova, A. A. Oxidative Stress Contributes to Lung Injury and Barrier Dysfunction Via Microtubule Destabilization. Am. J. Respir. Cell Mol. Biol. 2012, 47, 688−697. (56) Yang, S. K.; Xiao, L.; Li, J.; Liu, F.; Sun, L. Oxidative Stress, a Common Molecular Pathway for Kidney Disease: Role of the Redox Enzyme P66shc. Renal Failure 2014, 36, 313−320. (57) Aruna, G.; Ambika Devi, K. Oxidative Stress in Chronic Renal Failure. Int. J. Pharma Bio Sci. 2014, 5, B127−B133. (58) Leibowitz, A.; Volkov, A.; Voloshin, K.; Shemesh, C.; Barshack, I.; Grossman, E. Melatonin Prevents Kidney Injury in a High Salt Diet-Induced Hypertension Model by Decreasing Oxidative Stress. J. Pineal Res. 2016, 60, 48−54. (59) Galle, J. Oxidative Stress in Chronic Renal Failure. Nephrol., Dial., Transplant. 2001, 16, 2135−2137. (60) Araujo, A. M.; Reis, E. A.; Athanazio, D. A.; Ribeiro, G. S.; Hagan, J. E.; Araujo, G. C.; Damiao, A. O.; Couto, N. S.; Ko, A. I.; Noronha-Dutra, A.; Reis, M. G. Oxidative Stress Markers Correlate with Renal Dysfunction and Thrombocytopenia in Severe Leptospirosis. Am. J. Trop. Med. Hyg. 2014, 90, 719−723. (61) Sung, C. C.; Hsu, Y. C.; Chen, C. C.; Lin, Y. F.; Wu, C. C. Oxidative Stress and Nucleic Acid Oxidation in Patients with Chronic Kidney Disease. Oxid. Med. Cell. Longevity 2013, 2013, 301982. (62) Small, D. M.; Bennett, N. C.; Roy, S.; Gabrielli, B. G.; Johnson, D. W.; Gobe, G. C. Oxidative Stress and Cell Senescence Combine to Cause Maximal Renal Tubular Epithelial Cell Dysfunction and Loss in an in Vitro Model of Kidney Disease. Nephron Exp. Nephrol. 2012, 122, 123−130. (63) Ozbek, E. Induction of Oxidative Stress in Kidney. Int. J. Nephrol. 2012, 2012, 465897. (64) Massy, Z. A.; Stenvinkel, P.; Drueke, T. B. The Role of Oxidative Stress in Chronic Kidney Disease. Semin. Dial. 2009, 22, 405−408. (65) Sayre, L. M.; Perry, G.; Smith, M. A. Oxidative Stress and Neurotoxicity. Chem. Res. Toxicol. 2008, 21, 172−188. (66) Galano, A. Free Radicals Induced Oxidative Stress at a Molecular Level: The Current Status, Challenges and Perspectives of Computational Chemistry Based Protocols. J. Mex. Chem. Soc. 2017, 59, 231−262. (67) Wijtmans, M.; Pratt, D. A.; Valgimigli, L.; DiLabio, G. A.; Pedulli, G. F.; Porter, N. A. 6-Amino-3-Pyridinols: Towards Diffusion-Controlled Chain-Breaking Antioxidants. Angew. Chem., Int. Ed. 2003, 42, 4370−4373. (68) Wijtmans, M.; Pratt, D. A.; Brinkhorst, J.; Serwa, R.; Valgimigli, L.; Pedulli, G. F.; Porter, N. A. Synthesis and Reactivity of Some 6Substituted-2,4-Dimethyl-3-Pyridinols, a Novel Class of ChainBreaking Antioxidants. J. Org. Chem. 2004, 69, 9215−9223. (69) Kim, H. Y.; Pratt, D. A.; Seal, J. R.; Wijtmans, M.; Porter, N. A. Lipid-Soluble 3-Pyridinol Antioxidants Spare A-Tocopherol and Do Not Efficiently Mediate Peroxidation of Cholesterol Esters in Human Low-Density Lipoprotein. J. Med. Chem. 2005, 48, 6787−6789. (70) León-Carmona, J. R.; Alvarez-Idaboy, J. R.; Galano, A. On the Peroxyl Scavenging Activity of Hydroxycinnamic Acid Derivatives: Mechanisms, Kinetics, and Importance of the Acid-Base Equilibrium. Phys. Chem. Chem. Phys. 2012, 14, 12534−12543. J

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling (71) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. (72) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215−241. (73) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378−6396. (74) Kumar, M.; Busch, D. H.; Subramaniam, B.; Thompson, W. H. Role of Tunable Acid Catalysis in Decomposition of A-Hydroxyalkyl Hydroperoxides and Mechanistic Implications for Tropospheric Chemistry. J. Phys. Chem. A 2014, 118, 9701−9711. (75) Alarcón, P.; Bohn, B.; Zetzsch, C.; Rayez, M. T.; Rayez, J. C. Reversible Addition of the Oh Radical to P-Cymene in the Gas Phase: Multiple Adduct Formation. Part 2. Phys. Chem. Chem. Phys. 2014, 16, 17315−17326. (76) Denegri, B.; Matić, M.; Kronja, O. A Dft-Based Model for Calculating Solvolytic Reactivity. The Nucleofugality of Aliphatic Carboxylates in Terms of Nf Parameters. Org. Biomol. Chem. 2014, 12, 5698−5709. (77) Parkhomenko, D. A.; Edeleva, M. V.; Kiselev, V. G.; Bagryanskaya, E. G. Ph-Sensitive C-on Bond Homolysis of Alkoxyamines of Imidazoline Series: A Theoretical Study. J. Phys. Chem. B 2014, 118, 5542−5550. (78) Xie, H. B.; Li, C.; He, N.; Wang, C.; Zhang, S.; Chen, J. Atmospheric Chemical Reactions of Monoethanolamine Initiated by Oh Radical: Mechanistic and Kinetic Study. Environ. Sci. Technol. 2014, 48, 1700−1706. (79) Mishra, B. K.; Lily, M.; Chakrabartty, A. K.; Bhattacharjee, D.; Deka, R. C.; Chandra, A. K. Theoretical Investigation of Atmospheric Chemistry of Volatile Anaesthetic Sevoflurane: Reactions with the Oh Radicals and Atmospheric Fate of the Alkoxy Radical (Cf3)2chochfo: Thermal Decomposition Vs. Oxidation. New J. Chem. 2014, 38, 2813−2822. (80) Sen, K.; Mondal, B.; Pakhira, S.; Sahu, C.; Ghosh, D.; Das, A. K. Association Reaction between Sih3 and H2o2: A Computational Study of the Reaction Mechanism and Kinetics. Theor. Chem. Acc. 2013, 132, 1 DOI: 10.1007/s00214-013-1375-3. (81) Mandal, D.; Sahu, C.; Bagchi, S.; Das, A. K. Kinetics and Mechanism of the Tropospheric Oxidation of Vinyl Acetate Initiated by Oh Radical: A Theoretical Study. J. Phys. Chem. A 2013, 117, 3739−3750. (82) Da Silva, G. Reaction of Methacrolein with the Hydroxyl Radical in Air: Incorporation of Secondary O 2 Addition into the Macr + Oh Master Equation. J. Phys. Chem. A 2012, 116, 5317−5324. (83) Zhao, Y.; Truhlar, D. G. How Well Can New-Generation Density Functionals Describe the Energetics of Bond-Dissociation Reactions Producing Radicals? J. Phys. Chem. A 2008, 112, 1095− 1099.

(84) Galano, A.; Alvarez-Idaboy, J. R. Kinetics of Radical-Molecule Reactions in Aqueous Solution: A Benchmark Study of the Performance of Density Functional Methods. J. Comput. Chem. 2014, 35, 2019−2026. (85) Matsui, T.; Baba, T.; Kamiya, K.; Shigeta, Y. An Accurate Density Functional Theory Based Estimation of Pk a Values of Polar Residues Combined with Experimental Data: From Amino Acids to Minimal Proteins. Phys. Chem. Chem. Phys. 2012, 14, 4181−4187. (86) Baba, T.; Matsui, T.; Kamiya, K.; Nakano, M.; Shigeta, Y. A Density Functional Study on the Pka of Small Polyprotic Molecules. Int. J. Quantum Chem. 2014, 114, 1128−1134. (87) Galano, A.; Pérez-González, A.; Castañeda-Arriaga, R.; MuñozRugeles, L.; Mendoza-Sarmiento, G.; Romero-Silva, A.; Ibarra-Escutia, A.; Rebollar-Zepeda, A. M.; León-Carmona, J. R.; HernándezOlivares, M. A.; Alvarez-Idaboy, J. R. Empirically Fitted Parameters for Calculating Pka Values with Small Deviations from Experiments Using a Simple Computational Strategy. J. Chem. Inf. Model. 2016, 56, 1714−1724. (88) Galano, A.; Alvarez-Idaboy, J. R. A Computational Methodology for Accurate Predictions of Rate Constants in Solution: Application to the Assessment of Primary Antioxidant Activity. J. Comput. Chem. 2013, 34, 2430−2445. (89) Francisco-Marquez, M.; Galano, A. The Reactions of Plant Hormones with Reactive Oxygen Species: Chemical Insights at a Molecular Level. J. Mol. Model. 2018, 24, 255. (90) Galano, A.; Tan, D. X.; Reiter, R. J. On the Free Radical Scavenging Activities of Melatonin’s Metabolites, Afmk and Amk. J. Pineal Res. 2013, 54, 245−257. (91) Muñoz-Rugeles, L.; Galano, A.; Alvarez-Idaboy, J. R. The Role of Acid-Base Equilibria in Formal Hydrogen Transfer Reactions: Tryptophan Radical Repair by Uric Acid as a Paradigmatic Case. Phys. Chem. Chem. Phys. 2017, 19, 15296−15309. (92) De Grey, A. D. N. J. Ho2·: The Forgotten Radical. DNA Cell Biol. 2002, 21, 251−257. (93) Perez-Gonzalez, A.; Alvarez-Idaboy, J. R.; Galano, A. Dual Antioxidant/Pro-Oxidant Behavior of the Tryptophan Metabolite 3Hydroxyanthranilic Acid: A Theoretical Investigation on Reaction Mechanisms and Kinetics. New J. Chem. 2017, 41, 3829. (94) Eyring, H. The Activated Complex in Chemical Reactions. J. Chem. Phys. 1935, 3, 107. (95) Evans, M. G.; Polanyi, M. Some Applications of the Transition State Method to the Calculation of Reaction Velocities, Especially in Solution. Trans. Faraday Soc. 1935, 31, 875−894. (96) Truhlar, D. G.; Garrett, B. C.; Klippenstein, S. J. Current Status of Transition-State Theory. J. Phys. Chem. 1996, 100, 12771−12800. (97) Kuppermann, A.; Truhlar, D. G. Exact Tunneling Calculations. J. Am. Chem. Soc. 1971, 93, 1840−1851. (98) Marcus, R. A. Electron Transfer Reactions in Chemistry. Theory and Experiment. Pure Appl. Chem. 1997, 69, 13−29. (99) Collins, F. C.; Kimball, G. E. Diffusion-Controlled Reaction Rates. J. Colloid Sci. 1949, 4, 425−437. (100) Smoluchowski, M. Mathematical Theory of the Kinetics of the Coagulation of Colloidal Solutions. Z. Phys. Chem. 1917, 92, 129. (101) Einstein, A. Ü ber Einen Die Erzeugung Und Verwandlung Des Lichtes Betreffenden Heuristischen Gesichtspunkt. Ann. Phys. 1905, 322, 549. (102) Stokes, G. G. Mathematical and Physical Papers; Cambridge University Press: Cambridge, 1903; Vol. 3. (103) Galano, A.; Alvarez-Idaboy, J. R. Computational Strategies for Predicting Free Radical Scavengers’ Protection against Oxidative Stress: Where Are We and What Might Follow? Int. J. Quantum Chem. 2019, 119, No. e25665. (104) Galano, A.; Mazzone, G.; Alvarez-Diduk, R.; Marino, T.; Alvarez-Idaboy, J. R.; Russo, N. Food Antioxidants: Chemical Insights at the Molecular Level. Annu. Rev. Food Sci. Technol. 2016, 7, 335− 352. (105) Galano, A.; Francisco-Marquez, M. Reactions of Ooh Radical with B-Carotene, Lycopene, and Torulene: Hydrogen Atom Transfer K

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX

Article

Journal of Chemical Information and Modeling and Adduct Formation Mechanisms. J. Phys. Chem. B 2009, 113, 11338−11345. (106) Alberto, M. E.; Russo, N.; Grand, A.; Galano, A. A Physicochemical Examination of the Free Radical Scavenging Activity of Trolox: Mechanism, Kinetics and Influence of the Environment. Phys. Chem. Chem. Phys. 2013, 15, 4642−4650. (107) Marino, T.; Galano, A.; Russo, N. Radical Scavenging Ability of Gallic Acid toward Oh and Ooh Radicals. Reaction Mechanism and Rate Constants from the Density Functional Theory. J. Phys. Chem. B 2014, 118, 10380−10389. (108) Iuga, C.; Alvarez-Idaboy, J. R.; Russo, N. Antioxidant Activity of Trans -Resveratrol toward Hydroxyl and Hydroperoxyl Radicals: A Quantum Chemical and Computational Kinetics Study. J. Org. Chem. 2012, 77, 3868−3877. (109) Galano, A.; Alvarez-Idaboy, J. R. A Computational Methodology for Accurate Predictions of Rate Constants in Solution: Application to the Assessment of Primary Antioxidant Activity. J. Comput. Chem. 2013, 34, 2430−2445. (110) Villuendas-Rey, Y.; Alvarez-Idaboy, J. R.; Galano, A. Assessing the Protective Activity of a Recently Discovered Phenolic Compound against Oxidative Stress Using Computational Chemistry. J. Chem. Inf. Model. 2015, 55, 2552−2561. (111) Cordova-Gomez, M.; Galano, A.; Alvarez-Idaboy, J. R. Piceatannol, a Better Peroxyl Radical Scavenger Than Resveratrol. RSC Adv. 2013, 3, 20209−20218. (112) Lipinski, C. A.; Lombardo, F.; Dominy, B. W.; Feeney, P. J. Experimental and Computational Approaches to Estimate Solubility and Permeability in Drug Discovery and Development Settings. Adv. Drug Delivery Rev. 2001, 46, 3−26. (113) Ghose, A. K.; Viswanadhan, V. N.; Wendoloski, J. J. A Knowledge-Based Approach in Designing Combinatorial or Medicinal Chemistry Libraries for Drug Discovery. 1. A Qualitative and Quantitative Characterization of Known Drug Databases. J. Comb. Chem. 1999, 1, 55−68. (114) Veber, D. F.; Johnson, S. R.; Cheng, H. Y.; Smith, B. R.; Ward, K. W.; Kopple, K. D. Molecular Properties That Influence the Oral Bioavailability of Drug Candidates. J. Med. Chem. 2002, 45, 2615− 2623. (115) Leeson, P. D.; Davis, A. M. Time-Related Differences in the Physical Property Profiles of Oral Drugs. J. Med. Chem. 2004, 47, 6338−6348. (116) Zhu, H.; Tropsha, A.; Fourches, D.; Varnek, A.; Papa, E.; Gramatical, P.; Ö berg, T.; Dao, P.; Cherkasov, A.; Tetko, I. V. Combinatorial Qsar Modeling of Chemical Toxicants Tested against Tetrahymena Pyriformis. J. Chem. Inf. Model. 2008, 48, 766−784. (117) Boda, K.; Seidel, T.; Gasteiger, J. Structure and Reaction Based Evaluation of Synthetic Accessibility. J. Comput.-Aided Mol. Des. 2007, 21, 311−325.

L

DOI: 10.1021/acs.jcim.9b00517 J. Chem. Inf. Model. XXXX, XXX, XXX−XXX