Next Article in Journal
Evidence of the Disassembly of α-Cyclodextrin-octylamine Inclusion Compounds Conjugated to Gold Nanoparticles via Thermal and Photothermal Effects
Next Article in Special Issue
Fenofibrate Therapy Restores Antioxidant Protection and Improves Myocardial Insulin Resistance in a Rat Model of Metabolic Syndrome and Myocardial Ischemia: The Role of Angiotensin II
Previous Article in Journal
Characterisation of the Broadly-Specific O-Methyl-transferase JerF from the Late Stages of Jerangolid Biosynthesis
Previous Article in Special Issue
Curcumin Ameliorates Furazolidone-Induced DNA Damage and Apoptosis in Human Hepatocyte L02 Cells by Inhibiting ROS Production and Mitochondrial Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phenolic Melatonin-Related Compounds: Their Role as Chemical Protectors against Oxidative Stress

by
Annia Galano
1,*,
Romina Castañeda-Arriaga
1,
Adriana Pérez-González
2,
Dun-Xian Tan
3 and
Russel J. Reiter
3
1
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, Iztapalapa, 09340 Mexico City, Mexico
2
Consejo Nacional de Ciencia y Tecnología (CONACYT)—Departamento de Química, División de Ciencias Básicas e Ingeniería, Universidad Autónoma Metropolitana-Iztapalapa, Av. San Rafael Atlixco No. 186, Col. Vicentina, Iztapalapa, 09340 Mexico City, Mexico
3
Department of Cellular and Structural Biology, UT Health Science Center, San Antonio, TX 78229, USA
*
Author to whom correspondence should be addressed.
Molecules 2016, 21(11), 1442; https://doi.org/10.3390/molecules21111442
Submission received: 3 October 2016 / Revised: 17 October 2016 / Accepted: 24 October 2016 / Published: 29 October 2016
(This article belongs to the Special Issue Chemistry and Pharmacology of Modulators of Oxidative Stress)

Abstract

:
There is currently no doubt about the serious threat that oxidative stress (OS) poses to human health. Therefore, a crucial strategy to maintain a good health status is to identify molecules capable of offering protection against OS through chemical routes. Based on the known efficiency of the phenolic and melatonin (MLT) families of compounds as antioxidants, it is logical to assume that phenolic MLT-related compounds should be (at least) equally efficient. Unfortunately, they have been less investigated than phenols, MLT and its non-phenolic metabolites in this context. The evidence reviewed here strongly suggests that MLT phenolic derivatives can act as both primary and secondary antioxidants, exerting their protection through diverse chemical routes. They all seem to be better free radical scavengers than MLT and Trolox, while some of them also surpass ascorbic acid and resveratrol. However, there are still many aspects that deserve further investigations for this kind of compounds.

1. Introduction

Aerobic organisms are bound to the oxygen paradox, i.e., they cannot live without oxygen, but at the same time it represents a hazard to their health status [1]. The risk arises from the formation of oxidants, which is inherent to aerobic respiration. These species inflict structural damage to numerous molecules that are biologically important, including carbohydrates, lipids, proteins, and nucleic acids. Such damage is usually referred to as oxidative stress (OS), and can be potentiated by environmental and physiological factors that contribute to increase the oxidant amount. Some few examples are: pollution, radiation, consumption of certain drugs, cigarette smoke, heavy alcohol consumption, ischemia, infections, physical or mental stress, and aging [2,3,4,5,6,7,8,9,10,11,12,13,14,15].
OS has been held responsible—at least partially—for the onset and development of a wide spectrum of life threatening diseases like cancer [16,17,18], cardiovascular disorders [19,20,21], Parkinson’s and Alzheimer’s diseases [22,23,24,25,26,27], atherosclerosis [28,29,30], and diabetes [31,32,33]. Therefore, it is apparent that identifying molecules for protection against OS is a matter of vital importance. In addition to the enzymatic protection, there are many molecules that can offer chemical protection against OS. They are frequently referred to as antioxidants, and in the last decades have become the focus of numerous investigations.
Among the molecules that offer chemical protection against OS, melatonin (MLT, Figure 1) and related compounds stand out [34,35,36]. There are several reasons why MLT has been proven to be particularly efficient for that purpose. It has very low toxicity, even at rather large doses [37]. It can easily cross physiologic barriers because of its optimal size, partial solubility in water and high solubility in lipids [38,39]. After been metabolized, MLT protection against OS does not decrease. In fact it is maintained, or even increased, due to the antioxidant capacity (AOC) of its metabolites [40,41,42,43,44,45]. Moreover, it has been proposed that this family of compounds can act in a “task-division” way, with some members of the family being particularly efficient as free radical scavengers, and others mainly behaving as metal chelators [43]. This way of action promotes a wide-ranging protection against oxidants.
On the other hand phenolic compounds have also been identified as efficient protectors against OS [46,47,48,49,50,51,52]. Therefore, it is not surprising that some phenolic melatonin derivatives are very efficient for that purpose [43,53]. In fact, they seem to be an appealing set of molecules in the battle against OS. As a result; this review focusses on the information gathered so far on these compounds. Different aspects relevant to their chemical protection against OS are reviewed, including location and functionality. A variety of reaction mechanisms involved in their AOC is analyzed. The data reported so far is used to propose trends in activity, based on comparisons with other antioxidants. In addition, some perspectives and current challenges regarding the role of MLT, and its phenolic related compounds, as protectors against OS are discussed.

2. Oxidative Stress and Free Radicals

More than half a century ago, in a pioneer work on the subject, Gerschman and coworkers [54] proposed for the first time that free radicals (FR) are the toxic intermediates associated with oxygen poisoning and ionizing radiation. The most representative feature of FR is that they have one, or more, unpaired electrons. This peculiarity makes them highly reactive and, consequently, very harmful species. FR are able of triggering chain reactions, propagating the molecular damage distant from the initial site of attack.
However, FR are not intrinsically dangerous. On the contrary, at low to moderate concentrations they have important physiological roles. For example FR are involved in the cellular signaling [4,6,9] and defense [7,8] systems, as well as in the maturation of cellular structures [2], mitogenic responses [3,4,5,6], regulation of insulin receptor kinase activity [7], and in the apoptosis of defective cells [55,56]. The toxicity of FR, and the resulting OS, arises as a consequence of a chemical imbalance between their production and consumption [57], which increases the FR concentrations above healthy levels. Under such conditions, free radicals can become a serious hazard to human health. For example, OS has been associated with neurological disorders [16,58,59,60,61,62,63,64], cancer [65,66,67,68,69], diabetes [70,71], pregnancy disorders, fetal defects and pre-eclampsia [72,73,74,75], as well as with cardiovascular [19,20,71,76,77,78,79], pulmonary [80,81,82], renal [20,83,84], and ocular [85,86,87] diseases.
There are abundant sources that contribute to increase the FR amounts in living organisms, thus promoting the deleterious effects of OS. Endogenously-produced FR arise from infection, inflammation, ischemia, immune responses, aging, and mental or physical stress [88,89,90,91,92,93,94,95,96,97,98]. On the other hand, cigarette smoke, environmental pollution, certain drugs, heavy or transition metals, alcohol, and radiation constitute exogenous sources that induce FR formation [99,100,101,102,103,104,105,106,107,108,109,110,111,112,113]. The most abundant FR, in vivo, are—or result from—reactive oxygen species (ROS), reactive nitrogen species (RNS), and reactive sulfur species (RSS), in that order. The names and acronyms of some of them are provided in Table 1.
The FR chemical reactivity is directly associated with the damage that they can inflict to biological molecules, with OH being the most reactive and dangerous species. It can react very fast (at, or close to, diffusion-controlled rates) with a large variety of molecules [114]. Thus, OH can damage almost any chemical species in the vicinity of its formation site [115]. This radical is responsible for most of the ionizing radiation-induced tissue damage [116] and easily oxidizes DNA [117,118,119]. RO are less reactive than OH and more reactive than ROO, as long as R is the same—or similar—in both species [120,121,122,123,124]. In the context of OS, ROO species are particularly important. They are capable of diffusing to remote cellular locations, and are among the main products yielded by lipid peroxidation [125]. ROO are also among the FR that can be successfully scavenged to retard OS [126] because their half-lives are not too short, which allows antioxidants to efficiently deactivate them before they harm biological targets [127].
The hydroperoxyl radical (HOO) is the smallest of the ROO, and the conjugated acid of O2•−. The acid-base equilibria involving these two oxidants is crucial to OS since most of the damage attributed to the HOO/O2•− pair is actually inflicted by HOO, albeit it represents only about 0.3% of the HOO/O2•− pair in the cytosol of a typical cell [128]. It has been demonstrated that the reactions of unsaturated fatty acids with HOO are ~5 orders of magnitude faster than with O2•− [129], and that HOO is actually responsible for initiating fatty acid peroxidation [130]. In addition, HOO is more reactive than other ROO species when R is an alkyl or an alkenyl group [120,131], while this trend is reversed when R is a group with higher electron-withdrawing character (for example R = −CCl3) [131,132].
Regarding RNS, the direct toxicity of NO is expected to be minor due to its rather low chemical reactivity [133,134]. However, when reacting with O2•−, NO can produce a significant harmful species (ONOO) [135], which is capable of chemically damaging lipids, proteins and DNA [136,137,138], The reactivity of NO2 is in between those of NO and ONOO, which makes it a moderate oxidant [139]. On the other hand, RSS are yielded by the reactions of thiols with ROS and RNS [140], thus RSS are expected to be less reactive than the corresponding O and N species. However, this does not mean that they are not dangerous. On the contrary, it has been demonstrated that RSS are capable of damaging proteins [141,142,143].

3. Chemical Antioxidants

Chemical antioxidants are species that offer protection against OS by non-enzymatic ways of action, and have been proven to be helpful in the prevention and treatment of numerous OS-related health disorders [10,144,145,146,147,148,149,150,151]. They can be endogenously produced or acquired from foods, or diet supplements. Some examples of endogenous chemical antioxidants are melatonin, coenzyme Q10, glutathione, and lipoic acids; albeit some of them can also be involved in the enzymatic defense system. Exogenous chemical antioxidants come from numerous sources, which can be classified as natural or synthetic depending on how they are produced. Phenolic compounds, carotenes, tocopherols and ascorbic acid are a few examples of natural exogenous antioxidants. On the other hand, N-acetylcysteine, gallates and edaravone, are synthetically produced chemicals with important antioxidant activity.

3.1. Mechanisms of Action

Chemical antioxidants are diverse not only in their sources and structures, but also in the reaction mechanisms contributing the most to their protection against OS. This particular feature allows to differentiate between primary (chain breaking or Type I) and secondary (preventive or Type II) antioxidants [152]. Primary antioxidants prevent oxidation by scavenging FR, i.e., by directly reacting with them, yielding significantly less reactive species or terminating the radical chain reaction. On the contrary, secondary antioxidants retard oxidation by indirect means of action, i.e., by chemical routes that do not involve direct reactions with FR. For example, by metal chelation, decomposition of hydroperoxide into non-radical species, repairing primary antioxidants, deactivating singlet oxygen, or absorbing ultraviolet radiation. Some species, however, can behave as multiple-function antioxidants, i.e., exhibiting both primary and secondary AOC.
The most common reaction mechanisms involved in the primary AOC exerted by chemical antioxidants are schematically represented in Table 2. The RAF route involves the addition of the free radical to the antioxidant. Thus, the most important structural feature of the later is the presence of multiple bonds. Other important aspect regarding this route is that it may be limited by steric effects, and the radical’s electrophilicity. The more electrophilic the radical, the more likely it is involved in RAF reactions. On the other hand, SET reactions can occur through two different routes, depending on which species is the electron acceptor: the radical (SET-I) or the antioxidant (SET-II), with SET-I being common one.
Since PCET and HAT reactions yield exactly the same products, it is not easy to differentiate between them. In fact, when a global chemical process consists on an H transfer it is frequently assumed to be a HAT reaction, albeit it might actually correspond to PCET. In HAT reactions the proton and the electron are transferred together as a single entity, i.e., a hydrogen atom. On the contrary, in PCET reactions the electron and proton are transferred as two separated particles, but in a single step, without any stable intermediate. A common way of describing PCET is as a reaction with a proton and electron transferred from different sets of orbitals. Thus, while in HAT the donor and the acceptor are the same for both particles (the electron and the proton), in PCET they are different.
The SPLET and SEPT mechanisms both involve two elementary steps: one proton transfer and one electron transfer, but in the opposite order. The SPLET mechanism was first proposed by Litwinienko and Ingold [153] for the reactions of phenolic compounds. Currently, there is an overwhelming amount of evidence supporting the crucial role of this mechanism on the AOC of this family of compounds. The SEPT mechanisms is less common, albeit it is involved not only in the free radical scavenging processes but also in the oxidation of biological targets, such as the nucleosides in DNA [119].
The information regarding the reaction mechanisms involved in secondary AOC is less abundant than that already gathered for primary AOC. Metal chelation is particularly appealing when it inhibits the OH production, which—as mentioned before—is one of the most damaging oxidants in biological systems. Within the cells, OH is mainly produced by the Fenton reaction or the Haber-Weiss recombination (HWR). The latter can be globally written as:
O2•− + H2O23O2 + OH + OH
However, it only becomes physiologically important when it is catalyzed by metal ions [154], which actually corresponds to a two-steps reaction:
Mq+ + O2 → M(q−1) + 3O2
M(q−1) + H2O2 → Mq+ + OH + OH
The second step is the so-called Fenton reaction. The most likely metals involved in these reactions are iron and copper, which in biological media are mainly found as Fe(III) and Cu(II). Accordingly, the first step of the HWR is expected to play a crucial role in the OH production. That is why chelating agents able to inhibit the reduction of Fe(III) and Cu(II) may be effective in downgrading, or inhibiting, the OH production and the associated OS. In fact, metal chelation has been proposed as a therapy for Alzheimer’s disease [155].
There is more than one reaction mechanism that might be involved in metal chelation. The most frequently assumed is the direct chelation mechanism (DCM):
Mq+ + HnAtx → M(HnAtx)q+
While the coupled deprotonation-chelation mechanism (CDCM) might arise when the antioxidant has acid-base equilibria [156]:
Mq+ + HnAtx → M(HnmAtx)q−m + mH+
Metal chelation is also a crucial step in the OH-inactivating ligand (OIL) [157,158] behavior of chemical species, which can protect against OH-induced damage in two different ways [159]: (i) sequestering metal ions from reductants by inhibiting the reduction of metal ions; or (ii) deactivating OH after being produced via Fenton-like reactions. In the latter OH are still produced, but they are immediately scavenged by the OIL acting as a ligand in the metal chelate.
Secondary AOC by single oxygen quenching (SOQ) may involve different routes, including:
1O2 + Atx → 3O2 + Atx
1O2 + Atx → AtxO2
The first one is a singlet oxygen physical quenching, with the antioxidant remaining chemically unchanged; while the second actually corresponds to a singlet oxygen chemical quenching, i.e., a new chemical species is yielded. The physiological importance of antioxidants able to quench 1O2 has been related, for example, to the protection of the skin from exposure to ultraviolet-A radiation [160]. Secondary antioxidants can also exert their protection by repairing primary antioxidants and biomolecules. These repair processes usually involve hydrogen or electron transfer reactions.

Main Reaction Mechanisms Involved in AOC Assays

Currently there are numerous experimental assays that can be used to measure the antioxidant capacity (AOC) of chemical compounds. However, the main reaction mechanism governing the outcome of different AOC assays may not be the same. Therefore, it is important to know what the dominant chemical route is for each of them. This knowledge is crucial to select the most appropriate method, depending on the antioxidant to be tested, and also to properly interpret the obtained results. Some of the most widely used AOC assays are briefly summarized here, emphasizing on the features most relevant to the context of this review. The interested reader can find more detailed information on these—and other—AOC assays elsewhere [161,162,163,164,165,166].
Oxygen radical absorbance capacity (ORAC) assay: It measures the radical chain breaking ability of antioxidants by monitoring the inhibition of peroxyl radical-induced oxidation. This technique usually involves using Trolox as the antioxidant reference, and—as a result—the AOC of the tested antioxidants is reported as Trolox equivalents. ORAC essentially measures the hydrogen atom donating ability of antioxidants, thus it can be considered as an HAT-based method. This assay was originally developed for measuring hydrophilic antioxidants, but later it has adapted for both lipophilic and hydrophilic species [167].
2,2-Diphenyl-1-picrylhydrazyl (DPPH) assay: It is based on the neutralization of the DPPH radicals by antioxidants mainly by electron transfer, which implies that AOC is ruled by the electron donating capacity of the tested molecules. However, it has been suggested that the HAT route might also be involved as a marginal pathway [162,168]. Albeit the DPPH assay is among the most frequently used to obtain a first AOC evaluation, it has been argued that the results obtained with this method are not necessarily extrapolable to biological systems [169]. Moreover, its applicability for ranking antioxidants has been recently questioned [170].
Trolox equivalent antioxidant capacity (TEAC) assay: It measures AOC as the ability of antioxidants to scavenge the radical cation known as ABTS•+ (2,2′-azinobis(3-ethylbenzothiazoline-6-sulphonic acid)). Although it is usually classified as an ET-based method, the HAT mechanism can also contribute to the observed AOC. The relative importance of these two chemical routes is mostly determined by the structure of the antioxidant and the pH of the medium [162]. TEAC can be conducted in both hydrophilic and lipophilic media and is not significantly affected by the ionic strength of the medium [171]. The results obtained by this assay are frequently reported as Trolox equivalents.
Ferric reducing antioxidant power (FRAP) assay: It measures AOC as the capability of antioxidants to reduce Fe(III) or Fe(III)-ligand complexes, in acidic media. Thus, it is a typical electron transfer-based method. Therefore, it has been argued that it cannot be related with HAT-based AOC occurring in lipid systems but it can be used—in combination with other methods—for identifying the main reaction mechanisms involved in AOC [162]. Albeit the conventional FRAP assay was designed to test hydrophilic antioxidants, it was modified to enable simultaneous measuring the AOC of both hydrophilic and lipophilic molecules [172].
Theoretical chemistry can also be used to estimate AOC. A computational protocol has been recently developed to produce reliable quantitative data concerning the kinetics of radical-molecule reactions in solution. It is commonly referred to as the quantum mechanics-based test for overall free radical scavenging activity (QM-ORSA) [173], and has been extensively used to produce trends in primary AOC. There are some key points when using this protocol. It is required to use the same FR, and solvents of similar polarity, to assure the fairness in the comparisons; and calculation methods that are reliable for kinetics, such as LC-ωPBE, M06-2X, BMK, B2PLYP and M05-2X [174]. The QM-ORSA protocol has been validated by comparison with experimental results, and its uncertainties have been proven to be no larger than those arising from experiments [173]. It allows the use of two different scales: (i) the absolute, based on overall rate coefficients; and (ii) the relative, using Trolox as reference, and making separately analyses in aqueous and non-polar (lipid-like) media. In addition, it accounts for a wide variety of reaction mechanisms including SET, HAT, SPLET and RAF.

3.2. Melatonin (MLT)

MLT is an indoleamine with two side chains, a 5-methoxy group and 3-amide group. Its molecular weight is 232.2 g/mol, it has 17 heavy atoms and logP = 1.4. Accordingly, the MLT size, partial solubility in water and high solubility in lipids, promotes that MLT can easily cross physiologic barriers. Other relevant physicochemical properties of MLT are the number of hydrogen bond acceptors (4), bond donors (2), and rotatable bonds (4) as well as its molar refractivity (65.6) [53], topological polar surface area (54.12 Å2) [53] and pKa (12.3) [175]. According to the latter, it is expected that at physiological pH (7.4) neutral MLT is by far the dominant acid-base form of this compound.
On the other hand, MLT reduction potential has been reported to be 0.95 V [176], which indicates that it may interact with the respiratory complexes of the electron transport chain by donating and/or accepting electrons. This behavior is expected to increase electron flow, which is an effect that not all other antioxidants possess [177,178]. MLT exhibits an anodic wave at a maximum potential value of 0.73 V but not wave in the reverse scan [179], which was attributed to the possible instability of the oxidation products as a consequence of fast second order decay of the radicals. In addition, the presence of the anodic wave constitutes evidence of the electron donating ability of MLT and supports the electron donation hypothesis used to explain the scavenging action of this compound, first proposed by Hardeland et al. [180]. In this hypothesis it is assumed that MLT radical cation—yielded by electron donation—is stabilized by resonance, thus having a rather long live. In turn this species reacts with O2•− producing N1-acetyl-N2-formyl-5-methoxykynuramine (AFMK).
The evidence gathered so far on the AOC of MLT is so abundant, and compelling, that it has inspired the hypothesis that the primary function of MLT in living organisms is to protect them from OS [181]. MLT is capable of efficiently scavenging a wide variety of oxidants including OH [182], RO [183,184], CCl3OO [131], 1O2 [185,186], and NO [187,188]. Consequently, it reduces the molecular damage associated with FR generation in vivo [189]. MLT has also been found to inhibit lipid peroxidation [190,191,192,193,194]. In addition, the FR scavenging activity of MLT is shared by its metabolites, which allows that its AOC remains after MLT is metabolized. Such continuous protection is known as the antioxidant cascade [178,195,196], and makes MLT an exceptionally successful protector against OS, even at low concentrations.
It has been described that MLT exhibits protective effects against OS-related neurodegenerative disorders [197,198] and other OS-related diseases. For example, it was found to reduce the degenerative changes in experimental models of Alzheimer’s disease [199,200,201,202,203], while MLT administration to humans significantly slowed-down the progression of this disease [204,205]. MLT also has neuroprotective effect in Parkinson’s disease [206,207], and has been found to alleviate some of the secondary symptoms of cancer [208,209], and to be synergic with some radio- or chemo-therapies [210,211,212,213,214]. In addition, MLT reduces the toxic effects of some anti-cancer drugs like cisplatin and gossypol [182,215]. Moreover, it has even been determined that MLT may be effective in controlling metastatic breast cancer, both in vitro and in vivo, not only via inhibition of the proliferation of tumor cells but also through direct antagonism of a metastatic mechanism [216]. The antioxidant activity of MLT has been associated with its role in limiting ischemia-reperfusion injuries in the central nervous system (CNS) [217,218,219,220], heart [221], kidneys [222], liver [223,224,225], and lungs [226]. In addition, it has been hypothesized that melatonin may be synthesized in mitochondria [227], which would make MLT—and its metabolites—available for protecting mitochondria from OS [228].
On the other hand, MLT also increases the protective effects of glutathione, ascorbic acid and Trolox [229,230]. This was attributed to the capability of MLT to regenerate them by electron transfer processes. Repairing damaged biological targets is another pathway involved in secondary AOC. It has been found that MLT promotes the repair of oxidized DNA [179,231,232,233]. This is probably due to the MLT’s capability of transforming guanosine radical to guanosine by electron transfer [179]. As a secondary antioxidant, MLT has also been identified as efficient for counteracting the cytotoxic action of 1O2 [185] yielding AFMK [234].
MLT chelates aluminum, cadmium, lead, zinc, iron and copper [235] and drastically decreases the amounts of FR yielded by the interaction of Cu(II), Fe(II), Zn(II), Al(III) and Mn(II) with the β-amyloid peptide [236]. MLT also lowers the Cu(II)/H2O2-induced damage to proteins [237] and protects against copper-mediated lipid peroxidation [238], which led to the suggestion that the antioxidant and neuro-protective effects of melatonin may involve removing toxic metals from the CNS [238]. The protection of MLT against metal-catalyzed molecular damage was recently reviewed, and it was suggested that MLT may prevent the copper-induced FR generation in vivo by binding this metal [239]. Thus, according to the evidence gathered so far, it can be stated that MLT is an efficient multiple-function antioxidant.

3.3. Phenolic Compounds

Phenolic compounds are—arguably—the chemical family most widely investigated in the context of AOC. Their ability to protect against OS, and the associated health disorders has been profusely documented [46,48,240,241,242,243]. Like MLT, they are versatile antioxidants, able to deactivate diverse oxidants by various reaction mechanisms. Regarding their role as primary antioxidants, phenols have been found to scavenge FR mainly by HAT [244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272], PCET [273,274,275,276,277], and SPLET [278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298]; albeit, they can exert such protective action also by SEPT [299,300,301], RAF [302], SET-I [303,304] and SET-II [283,305]. As secondary antioxidants, phenolic compounds are capable of quenching 1O2 [306,307,308,309,310,311], repairing biomolecules [312,313,314,315] and chelating metals [316,317].
There is also abundant evidence on the beneficial health effects of phenolic compounds, which have been attributed to their antioxidant activity [48]. For example, they have been found to be effective in the prevention and therapy of skin disorders [318], to protect against cardiovascular [319,320,321,322] and neurodegenerative diseases [323,324], liver failure [325], and the toxic side effects of some pharmacological drugs [326]. In addition to their antioxidant related benefits, they have also shown anti-inflammatory [327] antimicrobial and antitumoral properties [50].
Phenolic compounds, particularly polyphenols are frequently present in the human diet. For example they are rather abundant in a wide variety of foods and beverages, including fruits, vegetables, wine, coffee and tea [328]. It has been reported that, after the intake of 10–100 mg of a particular phenolic compound, its maximum concentration in plasma rarely exceeds 1 mM [329]. However, it has also been assumed that the total phenol concentration in plasma is probably higher because of the presence of metabolites formed in body’s tissues or by the colonic microflora, which are still mostly unknown and not accounted for in the reported estimations [329]. At the same time, it seems that bioavailability can significantly change for different phenolic compounds, and that the most abundant phenols in our diet are not necessarily those with the best bioavailability profile [330].

4. Phenolic Melatonin-Related Compounds

Considering all the beneficial effects and the AOC of both MLT and phenolic compounds, it seems logical to assume that combining their structural features would lead to species with boosted or synergic activities. The phenolic MLT-related compounds analyzed in this review are shown in Figure 2.
Some of them are naturally-occurring molecules, while others were recently proposed as very promising antioxidants, based on computational-design strategies. The natural ones are involved in the tryptophan metabolic pathway in animals or plants (Figure 3). However, it is known that the metabolism of MLT is a highly complex process, involving both enzymatic and non-enzymatic (FR-induced) degradation [331]. The resulting products frequently overlap, making it difficult to identify which is the dominant degradation route, albeit under OS conditions the FR pathway is assumed to be the major one. In addition to what is shown in Figure 3, 2-hydroxymelatonin (2-HMLT) and 4-hydroxymelatonin (4-HMLT) are produced during the UV-induced metabolism of MLT [332]. 6-hydroxymelatonin (6-HMLT) is the primary hepatic metabolite of MLT in animals, while in plants 2-HMLT is the most abundant one [331,333]. This compound is also yielded by the oxidation of MLT by taurine chloramine [334].
The computationally-designed compounds shown in Figure 2 were chosen from a set of 19 melatonin analogues, intended to be better antioxidants than MLT, and to present both primary and secondary AOC. They were found to be among the best peroxyl radical scavengers identified so far, in aqueous solution, at physiological pH, and capable of downgrading OH production [53].

4.1. Location and Sources

The production of 5-hydroxytryptophan (5-HTP) and serotonin (5-HT) starts with dietary intake of L-tryptophan, which is present in a wide variety of foods including egg whites, chocolate, cod, dairy products, nuts and meats [335]. 5-HTP can also be taken as a dietary supplement, and it has been reported that it is well absorbed from consumed sources, with ~70% of the intake reaching the bloodstream [336]. In addition, the intestinal absorption of 5-HTP is not affected by the presence of other amino-acids, and it does not require a transport molecule. Accordingly, it may be taken with meals without decreasing it efficiency [337]. In addition, 5-HTP easily crosses the blood-brain barrier, thus it effectively contributes to increase the 5-HT synthesis in the CNS. The amount of endogenous 5-HTP available for this purpose depends on the availability of tryptophan, as well as on the activity of different enzymes; while the amount of 5-HTP that reaches the CNS is affected by how much 5-HTP is converted into 5-HT in the periphery [338].
5-HT is a ubiquitous molecule in nature. In plants it can be found in vegetables, fruits and nuts [339]. Some examples are cherries, coffee, tomatoes, Chinese cabbage, spinach, hot pepper, chicory, green onion, strawberry, lettuce, and rice [340,341,342,343]. In animals 5-HT is present in both vertebrates and invertebrates. In the particular case of humans, under normal conditions, 5-HT accounts for less than 2% of the L-tryptophan intake, which leads to a daily production of ~10 mg serotonin [339]. 5-HT is the neurotransmitter most widely distributed in the brain, albeit its amount in the CNS represents less than 5% of the whole body content [344]. It can be found in the pineal gland, serotonergic neurons, spinal cord and platelets; as well as in the liver, lungs, thyroid, bronchi, thymus and pancreas. However, the largest amount of 5-HT (about 80% of total content) is found in the gastrointestinal tract, particularly in the enterochromaffin cells [345], which are considered the site of synthesis and storage of 5-HT from tryptophan [346].
5-HT, NAS, 2-HMLT and 6-HMLT are present in cutaneous melatonin synthesis and metabolism with the latter being the main metabolite in epidermal cells [228]. On the other hand, while 5-HT is abundant in plants, N-acetylserotonin (NAS) is not [347,348,349,350]. However, it seems important to note that it is not only a precursor of MLT, but MLT can also be metabolized back into NAS [351]. Therefore, regardless of their relative abundance, it expected to find one of them wherever the other is present.
There is rather scarce information on whether 2-HMLT is enzymatically catabolized in vivo [196]. What most of the gathered evidence supports is that it is a product yielded from the reactions of MLT with chemical oxidants, such as the hypochlorous acid [352], oxoferryl hemoglobin [353], and OH [354]. The latter reaction also yields 4-HMLT. In addition, during the oxidation of cytochrome C by H2O2, in vitro, 2-HMLT is produced as an intermediate in the formation of AFMK [355]. It has also been observed that UV-B radiation-induced keratinocytes transform MLT into 2-HMLT [332]. This compound was found in rice seedlings, supporting its in vivo production from the MLT metabolism in plants [356]. On the other hand, 6-HMLT has been identified in the cerebral cortex, serum, heart, kidneys and liver of mice [205,357]. It has also been identified as the major MLT metabolite in the human skin [358,359]. Both, 2-HMLT and 4-HMLT were identified as products of the UV-induced metabolism of MLT in keratinocytes and cell-free systems [332].
The current information on the location and potential sources of 2-HMLT, 4-HMLT and 6-HMLT is less abundant than that concerning other phenolic MLT-related compounds. However, since they are produced by the interaction of MLT with oxidants—including ROS and RNS—it is expected that 2-HMLT, 4-HMLT and 6-HMLT are present in the same places as MLT, particularly under OS conditions. Since MLT is a ubiquitous molecule that can be found in plants and in many animal organs [357,359,360,361,362,363,364], its hydroxylated derivatives can be also formed in all of them. However, being as reactive towards oxidants as it will be demonstrated in following sections of this review, they are expected to have very short lives, which would make them hard to detect (particularly in excretion fluids, like urine).

4.2. Functions and Toxicity

Some of the species reviewed here are multifunctional molecules. In this section, their functions—other than AOC—are briefly described. It seems important to note than under OS conditions, just because of their AOC, the amounts of these compounds may be decreased affecting their other functions. This is particularly important for 5-HTP, 5-HT and NAS, since, apparently, the biological role of 2-HMLT, 4-HMLT and 6-HMLT is mainly to deactivate oxidants.
As mentioned before, 5-HTP is directly involved in the synthesis of 5-HT, while the latter and NAS are crucial to the MLT production. It has been shown that oral administration of 5-HTP can be effective in the treatment of depression [336,365]. It is assumed that 5-HTP supplementation normalizes the synthesis of 5-HT, a deficiency of which is believed to cause depression [366]. In addition to depression, 5-HTP has been found to be effective in the treatment of several conditions, including insomnia, fibromyalgia, cerebellar ataxia, binge eating and chronic headaches [337]. It has also been reported that 5-HTP supplementation inhibits leukocyte recruitment, serotonylation and allergic inflammation, which led to propose 5-HTP as a potential candidate in the treatment of allergy/asthma and the associated anxiety/depression symptoms [367]. 5-HTP was also found to be effective in reducing seizure-induced respiratory arrest, and was proposed as a possible therapy for preventing sudden unexpected death in epilepsy [368].
5-HT is also a multi-tasking molecule. It is involved in numerous physiological processes, such as peripheral and CNS neurotransmission, blood pressure regulation and smooth muscle contraction [339]. Probably its best known function is as a neurotransmitter–neuromodulator in the CNS. Its signaling pathways are involved in sensory processing, emotion regulation, cognitive control, autonomic control, and motor activity. Consequently, 5-HT modulates anxiety, fear, mood, stress, appetite, sleep, cognition, aggression and sexual behavior [344,369]. It is also a target of several physiological regulatory mechanisms and modulators such as gene transcription, psychotropic drugs, steroids and neurotrophic peptides [370]. In addition, it has been proposed that autism spectrum disorder, schizophrenia, bipolar disorder, impulsive behavior and attention deficit hyperactivity disorder, are all characterized by a dysfunction in the 5-HT pathway [371]. Deficient brain 5-HT synthesis, during development and adulthood, has been related to long-term neurodevelopmental disorders such as aggression, negative emotionality, and antisocial behavior [372]. Moreover, 5-HT has been identified to be concentrated in distinct brain regions collectively known as “the social-brain” [373,374].
In the periphery, 5-HT has been proposed to act as a pro-aggregator and vasoconstrictor when released from aggregating platelets, as an autocrine hormone when released from the enterochromaffin cells in the gut, pancreas and elsewhere; and as a neurotransmitter in the enteric plexuses of the gut [339]. In addition, 5-HT may be involved in the pathophysiology of hiccups based on its role in regulating the smooth muscles in the gastrointestinal tract to increase the tone and facilitate peristalsis, its vasoconstrictor properties, and its ability to dilate blood vessels of the heart and skeletal muscles [346]. 5-HT has also been identified as an endocrine hormone, a paracrine factor, or a growth factor; involved in mucosal growth/maintenance, gastrointestinal motility, intestinal inflammation, enteric neurogenesis, hepatic regeneration and osteogenesis [345].
At the same time, at elevated levels, 5-HT becomes toxic, leading to what is commonly referred to as the serotonin syndrome (SS). Its symptoms are numerous and affect the cardiovascular, gastrointestinal, autonomic, muscular, and central nervous systems [335]. Some examples are hypertension, disorientation, dizziness, flushing, hyperthermia, hyperreflexia, and myoclonus. In addition, elevated levels of 5-HT in blood—which are associated to disruption of the 5-HT/NAS/MLT pathway—have been identified as the most common marker in autism spectrum disorders [375].
To minimize the risk of SS it has been recommended not to administer 5-HTP in combination with serotonergic antidepressants [337,338]. There has been some cases of SS in patients concurrently taken L-tryptophan and fluoxetine, or switching from one serotonin reuptake inhibitor to another. On the contrary, there are no reports associating SS with the consumption of 5-HTP in monotherapies or combined with other medications. The most frequent adverse effects of 5-HTP identified so far involve the gastrointestinal tract, and include vomit, nausea and diarrhea; albeit insomnia, headache, and palpitations have also been observed [338]. While intravenous administration of 200–300 mg of 5-HTP can induce memory impairment, confusion and anxiety; these effects seldom appear with oral administration, especially at lower doses. For example, at dosages lower than 50 mg/kg/day no toxic effects have been found in connection with 5-HTP administration [338].
NAS, like MLT, functions as a signal molecule triggering plant defense responses [350,376] and growth [377]. In animals, NAS and 6-HMLT protect keratinocytes against UVB-induced OS and DNA damage [378]. They also exhibit membrane stabilizing activity in liver injury models [379]. NAS (and also MLT) improves membranes fluidity under OS conditions [380,381], which led to presume that they stabilize cellular membranes by preventing FR-induced lipid peroxidation [228,380,382]. In addition, the potential role of NAS, and melatonin, in the treatment of multiple sclerosis has been recently highlighted [383]. It has also been found that NAS can protect against acute hepatic ischemia-reperfusion [384].

4.3. Antioxidant Activity

The combined AOC of MLT and its metabolites may be responsible for the melatonin’s ability of deactivating several equivalents of oxidants. It is an intricate process that can be characterized as multifunctional [385] involving not only FR scavenging activity but also secondary AOC. For example their role as metal chelators may be highly beneficial for reducing the associated OS, since it would lead to the inhibition of OH [385]. It has been proposed that MLT and its metabolites may act in a “task-division” way, with some of them being particularly efficient as FR scavengers, and others mainly behaving as metal chelators [43]. Since scavenging FR and other oxidants result in a decrease of the associated damaging events, it would also imply an inhibition of protein oxidation, lipid peroxidation, mitochondrial damage and DNA destruction. Therefore, the antioxidant protection exerted by MLT, and related compounds, is expected to be efficient in maintaining a healthy redox status.

4.3.1. 5-Hydroxytryptophan (5-HTP)

It has been shown that 5-HTP is efficient in inhibiting OS-induced damage. For example, its beneficial effects in inflammatory diseases have been attributed to its AOC [386]. 5-HTP also inhibits iron-induced lipid peroxidation processes [387] and prevents the oxidation of proteins and lipids, induced by iron-ascorbate mixtures [388]. These effects may arise from the ability of 5-HTP to chelate metal ions [389]. 5-HTP is also capable of suppressing UV-induced apoptosis in human monocytes [390], to inhibit the oxidative damage induced by tert-butylhydroperoxide (t-BuOOH) on human fibroblast [391], to preserve membranes fluidity under OS conditions [392] and to suppresses inflammation and collagen-induced arthritis by decreasing the production of pro-inflammatory mediators [393]. Based on in vitro investigations, it has been proposed that 5-HTP is more potent as an OH scavenger than MLT and ascorbic acid [394] and Trolox [395]. In addition 5-HTP has been described as efficient for scavenging α,α-diphenyl-β-picrylhydrazyl (DPPH) and OH radicals, and to cause about 95% inhibition of linoleic acid peroxidation [396]. 5-HTP also shows NO scavenging activity [397] and AOC protection on hyperglycemia-induced oxidative stress [398], albeit in both cases it was described as less efficient than MLT.

4.3.2. Serotonin (5-HT)

It has been proposed that 5-HT protects membranes from lipid peroxidation through its FR scavenging activity, which is mainly exerted in the aqueous phase, or at the water-membrane interface [399]. Such protective activity has been held responsible for the attenuated secondary tissue damage in the CNS when serotonin is released at sites of brain damage or inflammation [400]. The antioxidant activity of 5-HT was also demonstrated by its suppressive effects on phagocytosis-associated, luminol-enhanced chemi-luminescence [401]. Its action was rationalized in terms of its reactions with ROS, found to be dose-dependent, and led to the proposal that 5-HT may modulate various aspects of cell-mediated defense reactions.
The in vitro antioxidant and FR scavenging activities of 5-HT have been evaluated using different methodologies and butylated hydroxyanisole (BHA), butylated hydroxytoluene (BHT), α-tocopherol and Trolox as reference compounds [402]. It was found that 5-HT is capable of fully inhibiting lipid peroxidation of a linoleic acid emulsion, being more efficient for that purpose than the references. 5-HT was also found to effectively scavenge DPPH, 2,2′-azino-bis(3-ethylbenzo-thiazoline-6-sulfonic acid) (ABTS•+), and hydrogen peroxide (H2O2) [402] as well as N,N-dimethyl-p-phenylenediamine (DMPD) radical [403]. The activity of 5-HT for the latter was found to be higher than that of MLT, which was attributed to its phenolic group. In addition, it was proposed that 5-HT, by scavenging peroxidase-derived ROS, may protect human natural-killer cells from oxidative damage at inflammatory sites [404].
The AOC of 5-HT on neuronal tissues have been examined by studying the oxidative damages of brain synaptosomal components induced by Fe(II) and ascorbate [405]. The oxidation processes were efficiently attenuated by 5-HT, which was also capable of significantly decreasing OH production and restoring the synaptosomal Ca(II) uptake, lowered by Fe(II) and ascorbate. In addition, it has been reported that 5-HT also exhibits ferric ion reducing power and ferrous ion chelating activities [402]. Accordingly, it seems that 5-HT plays protective roles against OS by directly scavenging FR, and also by sequestering metals and inhibiting FR production.

4.3.3. N-Acetylserotonin (NAS)

Although NAS is both a precursor and a metabolite of MLT, its protective effects against OS seem to be independent from those of MLT [380,406]. In fact, it was hypothesized that since about 15% of melatonin is demethylated into NAS in vivo, some of the beneficial effects of MLT might be attributed to NAS. In addition, it has been proposed that the antioxidant effects of NAS against ROS, induced by t-butylated hydroperoxide and diamide, are higher than those of MLT [407].
It has been reported that NAS offers neuroprotection by inhibiting autophagy activation and mitochondrial death pathways [408] and to protect from H2O2-induced OS injuries [409]. NAS also defends against 6-hydroxydopamine-induced neurotoxicity [410]. Moreover, it has been suggested that the neuroprotective properties of NAS—and MLT—might mediate their cognition-enhancing effects [411]. NAS also exhibits antioxidant and anti-aging activities [412], as well as protective effects against β-amyloid-induced neurotoxicity [411] and light-induced degeneration of retinal photoreceptor cells [413]. There is also evidence on the NAS role in inhibiting H2O2-induced cell death and ROS production in hepatocytes cells [409]. NAS also reduces lipid peroxidation caused by iron, H2O2 and 2,2′-azobis (2-amidinopropane), also known as AAPH [380,406,414,415]. It was found that NAS administration to mice resulted in decreased lipid peroxidation and increased glutathione peroxidase in brain and kidney [412]. It was also found that NAS suppressed the glutamate-induced lipid peroxidation in retinal homogenates [416]. The effects of NAS in reducing lipid peroxidation is responsible for the protective role played by NAS in preserving optimal fluidity of the biological membranes [381].
NAS inhibits copper-induced lipids oxidation [417,418,419] and Cr(III)-induced DNA oxidative injuries [420]. It also has protective effects against Fe(II)-ascorbate dependent lipid peroxidation [421], iron-induced lipid peroxidation and lipids autoxidation [422]. It has been suggested that the antioxidant effects of NAS might reinforce its anti-aging, cognition-enhancing, antihypertensive, antidepressant, and antitumor effects [423]. It was also proposed that NAS—and its derivatives—might be useful in protecting against OS-related disorders such as cell death and mutagenesis as well as diseases like cancer, sepsis, post-ischemic trauma, and Parkinson’s and Alzheimer’s diseases. NAS was also suggested to be directly involved in the self-protective antioxidant system of the retina, as an efficient physiological FR scavenger within the photoreceptor cell [406].

4.3.4. 6-Hydroxymelatonin (6-HMLT)

Not only 6-HMLT has been identified as an effective protector against OS-induced molecular damage, but also it has been proposed (together with NAS) as responsible for some—or much—of the AOC of MLT in vivo [228]. It was also suggested that since NAS and 6-HMLT are more hydrophilic than melatonin, their FR scavenging actions are exerted mainly in the aqueous phase, or at the water-lipid interface, while MLT positions itself within the lipid bilayer where it protects membrane proteins against FR attacks [228].
6-HMLT has been found to inhibit thiobarbituric acid-induced lipoperoxidation, which is attributed to ROS [424]. It was also suggested that the AOC of 6-HMLT is even larger than that of MLT in this context, which might be because of the presence of a hydroxyl group in the benzene ring, i.e., a phenol moiety. 6-HMLT was found to protect against cyanide-induced OS, which was attributed to its ability to downgrade KCN-induced O2•− generation [425]. 6-HMLT is also capable of protecting rat brain homogenates against iron-induced lipid peroxidation, in vitro [426], and against quinolinic-acid-induced oxidative neurotoxicity in the rat hippocampus by efficiently scavenging ROS [427]. It has been demonstrated that 6-HMLT (but also MLT, 5-HT and 5-HTP) is capable to quenching 1O2 at rather fast rates [186]. This finding led to the proposal that 6-HMLT should present neuro-protective effects [427].
Regarding its secondary AOC, it has been reported that 6-HMLT efficiently protects against oxidative damage induced by UV irradiation [428]. Moreover, 6-HMLT was proposed as partially responsible for the potential benefits of incorporating MLT into sunscreens. In addition, there is evidence that this compound can reduce Fe(II)-induced lipid peroxidation and necrotic cell damage in the rat hippocampus in vivo [429]. Accordingly, it seems that 6-HMLT exerts both primary and secondary AOC, albeit it has been characterized as a better primary antioxidant [43]. On the contrary, it was suggested that, in the presence of metal ions, 6-HMLT might induce DNA damage via non-o-quinone type of redox cycle leading to carcinogenic effects [97]. Therefore, despite of the fact that most of the investigations reported so far—on the OS related effects of 6-HMLT—indicate that it is beneficial, further studies aiming to provide more information on this subject are still desirable.
To the best of our knowledge, there are no reports yet on the independent AOC of 2-HMLT and 4-HMLT. They are mainly identified as oxidation products of MLT. However, they are structurally similar to NAS and 6-HMLT, considering that they all are mono-phenolic derivatives of MLT. Accordingly, it might be expected that they may also have protective effects against OS by scavenging FR or by any of the chemical routes involved in secondary AOC. These two compounds, definitively deserve further investigations regarding their potential roles as antioxidants.

4.3.5. Computationally-Designed Molecules

Albeit numerous synthetic MLT derivatives have been obtained in the last years, to our best knowledge none of them have a phenolic group. However, very recently a series of computationally-designed molecules presenting this feature were investigated [53]. For that purpose the Density Functional Theory (DFT) was used. The calculations were carried out with the 6-311+G(d,p) basis set and the continuum solvation model based on density (SMD) [430] using water as solvent. The M05-2X and M05 functionals [431] were used for geometry optimizations and frequency calculations for the systems without and with copper, respectively. Thermochemical and kinetic information on the AOC of the designed compounds was obtained and use to identify the most promising molecules.
The primary AOC was evaluated using the reactions with HOO, and all the modeled compounds were found to be much better peroxyl radical scavengers than MLT. All the molecules shown in Figure 2 were predicted to react with HOO faster than NAS, and at rates similar to those of the reactions with 6-HMLT. This is relevant since NAS and 6-HMLT were previously identified as particularly efficient for that purpose [43]. Moreover, compounds C1, C2 and C3 were predicted to react with the target radical at diffusion-limited rates, which makes them excellent candidates to be efficient as primary antioxidants. In addition, it has been previously reported that phenolic compounds can be regenerated under physiological conditions. Such regeneration takes place in such a way that these compounds can scavenge several radical equivalents in the process, two per cycle (one HOO and one O2) [257,260,261,263,272,432,433,434]. Accordingly, the modeled compounds are expected to be capable of deactivating several FR equivalents.
The secondary AOC was evaluated as the Cu(II) chelation ability, and the inhibition of OH production. The idea was that while Cu(I) is required for producing OH, Cu(II) is the most abundant and stable oxidative state of copper. Therefore, chelating agents capable of decreasing the viability of Cu(II) reduction should be effective for preventing, or inhibiting, OH production, via the Fenton reaction, and the consequential OS. Compounds C1, C2 and C3 were identified as capable of turning off the Cu(II) reduction induced by O2•− and Asc, thus inhibiting the associated OH production. C1 and C2 were identified as the species with the best multifunctional AOC, and both fulfill the Lipinski’s [435] and Ghose’s [436] rules for orally active drugs. Thus, it is assumed that they will not present problems of bioavailability, poor permeation or absorption. However, C1 was the one proposed as the best prospect for possible application based on potential toxicity and synthetic accessibility estimations. It is expected that computationally-designed antioxidants with promising estimated properties would be actually synthesized in the near future, so their protective effects against OS-related injuries can be tested.

4.3.6. Chemical Pathways

The relative importance of the different reaction mechanism that may be involved in the AOC of chemical compounds is probably one of the least explored aspects of this area of investigation. Regarding the compounds reviewed here, more investigations in the subject are still needed, albeit some advances had been made. Hydrogen transfer has been identified as a major pathway for the reactions of 5-HTP and 5-HT with DPPH [437], and the HOO scavenging activity of NAS, and 6-HMLT, in lipid media [43]. However, it is important to note that no investigations have been made on whether they are actually HAT or PCET reactions.
The RAF mechanism was proposed as relevant for the reactions of α-hydroxyethyl radicals with 5-HTP, 5-HT and MLT [438], while the chemical route referred here as SET-I has been identified as important for the primary AOC of 5-HTP and 5-HT [437,438,439]. On the other hand, the SPLET mechanism has been characterized as the thermodynamically preferred antioxidant mechanism in water for MLT and 60 meta- and ortho-substituted MLT derivatives [440]. It was also identified as viable for 5-HTP when scavenging a wide variety of free radicals [441] and for the reactions of NAS and 6-HMLT with HOO when they take place in aqueous solution, at physiological pH [442]. Under the same conditions, the SPLET mechanism was proposed as the main chemical route for the HOO scavenging activity of the computationally-designed molecules shown in Figure 2 [53].
Regarding secondary AOC, metal chelation has been proven to be important for MLT and related compounds [238,239,385,402]. It has been proposed that NAS and 6-HMLT are capable of turning off the OH production induced by copper-ascorbate mixtures and partially inhibiting the HWR [43]. Therefore, NAS and 6-HMLT are predicted to behave mainly as OIL type (ii). On the other hand, the computationally-designed phenolic MLT derivatives C1, C2 and C3 were characterized as multi-functional antioxidants; while C4 and C5 were designed to be primary antioxidants. For all of them, SPLET was identified as the main reaction mechanism involved in their primary AOC.

4.4. Comparisons with Other Antioxidants

Trends in AOC are crucial to design efficient strategies against OS, since they allow identifying the most efficient compounds for that purpose. However, it is a difficult task since the available data was not necessarily acquired using the same assay, or under the exact same conditions. For example, it has been reported that the AOC trends for the same set of antioxidants can be different in oil/water emulsion than in bulk corn oil, and also depending on the pH and the AOC assay [443,444]. Prior et al. [162] have called attention to the current lack of a universal AOC assay, and according to Frankel and Meyer “There is a great need to standardize antioxidant testing to minimize the present chaos in the methodologies used to evaluate antioxidants” [163].
Despite of all the challenges involved in making fair comparisons among antioxidants, and the rather scarce amount of information on the subject for phenolic MLT-related compounds, some trends have been established using both experimental and theoretical approaches. According to the experimental evidence, NAS was identified as more effective than MLT—but similar to Trolox—as a ROO scavenger [407] and for protecting against iron and lipopolysaccharide-induced lipid peroxidation [412]. The same trend was reported for lipid autoxidation and iron-induced lipid peroxidation [421,422], the copper-mediated oxidation of low density lipoproteins [417], and the Fe(II)-ascorbate induced lipid peroxidation in the bovine retina [421].
Regarding the other phenolic compounds reviewed here, 6-HMLT was found to be more efficient than MLT for inhibiting the lipo-peroxidation induced by thiobarbituric acid, which is attributed to the production of ROS [424]. 5-HTP was described as a more potent OH scavenger than MLT [394,445] and also for scavenging DMPD [129]. The free radical scavenging activity of 5-HT was found to exceed that of BHA, BHT, α-tocopherol and Trolox [130]. In addition, 5-HTP and NAS were characterized as better antioxidants than BHT in a triglyceride-system, albeit they were ineffective in a liposome-system, while MLT is unable of protecting polyunsaturated fatty acids (PUFAs) against lipid peroxidation in both cases [446]. These findings support the idea that evaluating AOC depends on the assay system.
On the other hand, the QM-ORSA protocol was designed to obtain kinetic data based on the antioxidant definition by Halliwell and coworkers [447,448], i.e., “any substance that when present at low concentrations compared to that of an oxidizable substrate would significantly delay or prevent oxidation of that substrate”. Therefore, for a molecule to be successful as an antioxidant, it should react faster than the species needing protection, which implies that rate constants (k) would be an optimum criterion for making AOC comparisons. The most abundant data obtained so far using QM-ORSA, correspond to the HOO scavenging activity of antioxidants. Based on the results obtained with QM-ORSA, NAS and 6-HMLT have been proposed as more efficient for scavenging ROO than MLT [43] which is in line with the experimental evidence. In addition, the computationally-designed MLT derivatives shown in Figure 2 were also predicted to be better peroxyl scavengers than the parent molecule [53].
There are also experimentally-measured rate constants for the reactions between FR and the compounds reviewed here. They are reported in Table 3, where the label theoretical means that the k values were obtained with the QM-ORSA protocol. These values were estimated considering HOO/O2•− equilibrium to obtain rate constants in line with experimental measurements. HOO has a pKa equal to 4.8, which means that in aqueous solution, at pH = 7.4, its molar fraction is only 0.0025. Since the most abundant data correspond to the reactions with HOO, a plot showing the trends in activity for scavenging this radical has been constructed to facilitate comparisons (Figure 4). In addition to the molecules of interest, the available data of other antioxidants, some of which are frequently used as reference, has also been included in this figure for comparison purposes.
According to the gathered kinetic data, MLT and its phenolic-related compounds all react at diffusion-limited rates with OH both in non-polar and polar solvents, i.e., they all are excellent OH scavengers. However, it is important to insist on the fact that this radical is highly reactive towards a wide variety of chemicals, despite the medium in which the reactions take place [457,458]. Thus, it is expected to react with almost any molecule in the vicinity of its formation site [115], which implies—arguably—that scavenging OH is not the best strategy to reduce OS, but to prevent its formation.
Despite the fact that the gathered kinetic data regarding NO2 and O2•− is rather scarce, some trends can be established (Table 3). It seems that the pH has an important influence on the rate at which NO2 is scavenged by the analyzed compounds. For example, it has been reported that the rate constant of its reaction with 5-HTP increases from 9.0 × 105 to 5.6 × 107 M−1·s−1, as the pH goes from 5 to 9 [450]. This can be rationalized considering that 5-HTP is involved in acid-base equilibria, thus as the pH increases so do the molar fractions of the deprotonated species. Accordingly, it seems that the AOC of 5-HTP increases with its deprotonation degree. On the other hand, the k value for the MLT + NO2 reaction was estimated to be 3.7 × 106 M−1·s−1, at pH = 7. Since the pKa of MLT is 12.3 [175], the fraction of its deprotonated species is negligible at neutral pH, and at any pH of physiological interest (3 ≤ pH ≤ 10). This means, that under such conditions, the reactivity of MLT towards NO2 should be rather independent of the pH. In addition, given the trend found for the reaction of 5-HTP with this radical at pHs equal to 5 and 9, it could be expected a rate constant in the order of 106 M−1·s−1, at pH = 7. Accordingly, it seems fair to propose that MLT and 5-HTP should have similar NO2 scavenging activity at physiological pH (pH = 7.4). For the reactions with O2, the k values suggest that the reactivity trend is MLT < 5-HTP < 6-HMLT, but with only small differences in reactivity. In addition, since the k value for MLT (Table 3) was measured using water as solvent (presumably at acid pH), while for the other two the solvent was not specified, this trend needs further confirmation.
The data involving HOO is the most abundant, and probably the most adequate to make comparisons, since all the k values analyzed here were obtained in a similar way, i.e., using the QM-ORSA protocol, and at the same pH for the reactions in aqueous solution. The trends in primary AOC, based on the reactions with HOO, is predicted to be 4-HMLT > NAS > caffeic acid > propyl gallate > resveratrol > ascorbic acid ≈ 6-HMLT > Trolox > MLT > caffeine, in non-polar (lipid) solution. Please note that the rate constants of C1-C5 were not included in the comparison because they have not been estimated yet in this media. In aqueous solution, at physiological pH, the primary AOC trend changes to C2 ≈ 4-HMLT ≈ C1 > C3 > propyl gallate > caffeic acid > ascorbic acid ≈ 6-HMLT > C5 > C4 ≈ resveratrol > NAS > Trolox > MLT > caffeine.
These trends strongly suggest that all the phenolic MLT-related compounds are better peroxyl radical scavengers than the parent molecule and Trolox, regardless of the polarity of the environment. This indicates that the presence of a phenolic moiety increases the FR scavenging activity within the MLT family. 4-HMLT was found to be in the top of the list in both cases, aqueous and lipid solution. The k values for 6-HMLT indicate that its primary AOC is similar to that of ascorbic acid but lower than that of propyl gallate and caffeic acid, also in both media. On the contrary, its relative activity with respect to resveratrol changes depending on the solvent’s polarity, i.e., it is lower in lipid and higher in water. Regarding the computationally-designed molecules, C2 and C1 seem to be the most promising ones as FR scavengers and as efficient as 4-HMLT, closely followed by C3. Moreover, they are among the best peroxyl radical scavengers reported so far. Their reactions with peroxyl radicals are predicted to be faster than those of any of the naturally occurring phenolic MLT-derivatives analyzed here, except 4-HMLT. Conversely, C4 and C5 are less efficient than 6-HMLT.
In addition, considering the above proposed trends, which identify 4-HMLT as the MLT metabolite with the highest primary AOC, a daring question arises: Is it actually 4-HMLT a minor metabolite of MLT (compared to 6-HMLT in animals and 2-HMLT in plants) or it is produced in amounts higher than previously thought but rapidly consumed because of its high reactivity towards FR? This probably deserves further investigations. In any case, what it seems to be clear is that 4-HMLT is a very efficient peroxyl radical scavenger, with higher activity than several widely recognized reference antioxidants.
Since the most abundant data gathered so far, concerning secondary AOC, involve metal chelation, it would be the one analyzed here. It has been found that MLT has a higher Fe(II) chelating ability than BHT, BHA and α-tocopherol [459], while NAS is more efficient than MLT for protecting against iron-induced lipid peroxidation [421,422]. NAS was also reported to be a better protector than MLT against copper-mediated oxidation of low density lipoproteins [417]. In addition, like their parent molecule, NAS and 6-HMLT were predicted to be capable of chelating Cu(II) [43,45,385]. However, although they were found to fully inhibit the OH production induced by Cu(II)-ascorbate mixtures, their inhibitory effects on the HWR were predicted to be only partial. Thus, the secondary AOC of MLT may be better than those of NAS and 6-HMLT.
Regarding the computationally-designed MLT phenolic derivatives, C1-C3 were identified as efficient for chelating Cu(II) [53]. Moreover, for Cu(II) chelates with them as ligands, the reactions with both O2•− and Asc become endergonic and dramatically slower than those involving free copper. Accordingly, they were predicted as efficient for inhibiting OH production, under physiological conditions via HWR. Considering that they also are excellent for scavenging FR, these compounds were proposed as promising multifunctional antioxidants.
Based on the previously analyzed data, a rough classification can be done regarding the AOC of the investigated compounds. It is provided in Figure 5, where MLT and some of its non-phenolic metabolites have been included for comparison purposes.
2-HMLT was not included in this figure because there is not enough quantitative data on the AOC of this compound. Hopefully, future investigations on this subject will provide more information on the AOC of 2-HMLT. In addition, the place of 5-HTP and 4-HMLT probably needs further confirmation. This has been indicated in the figure with a question mark next to their acronyms. It seems worthwhile to clarify that this classification was made based on current evidence, and considering the main kind of chemical protection exerted by the analyzed compounds against OS. However, it does not ruled out secondary AOC for a compound classified as primary, or vice versa. According to the gathered data, it seems that most of the phenolic MLT-related compounds are capable of counteracting OS as both, primary and secondary antioxidants, i.e., they present multi-functional AOC. This is a very desirable characteristic due to the wide variety of oxidants that are present in biological systems. This also means that they are not only able of scavenging FR, after they are produced, but also to inhibit their production under physiological conditions. Moreover, being as promising protectors against OS as they seem to be, they deserve further investigations that help to gain a deeper understanding on their chemical mechanisms of action.

5. Concluding Remarks

There is no doubt that OS represents a serious hazard to human health, thus molecules capable of offering protection against OS through chemical routes—in addition to the enzymatic pathways—are crucial to maintain a good health status. The phenolic and MLT families of compounds have both been identified as very efficient for that purpose. Accordingly, it is only logical to expect that phenolic MLT-related compounds are (at least) equally efficient as antioxidants. However, these compounds have been less investigated than phenols, MLT and its non-phenolic metabolites in the context of AOC.
The evidence gathered so far strongly indicates that MLT phenolic derivatives can act as both primary and secondary antioxidants. It also indicate that they may exert their protection against OS through diverse chemical routes including HAT, SPLET, metal chelation and inhibition of the OH production. All of the MLT-related compounds reviewed here seem to be better FR scavengers than MLT and Trolox, and most of them also surpass resveratrol, ascorbic acid, propyl gallate and caffeic acid.
On the other hand, OS and AOC are complex and manifold processes. They may involve multiple chemical species and reaction pathways, and may be influenced by several environmental factors such as the polarity of the media, the pH in aqueous solution, and the presence of other chemical species. There are many of these aspects that still deserve further investigations, particularly for the kind of compounds reviewed here. For example, very little is known regarding the potential AOC of 2-HMLT, or on the possible role of 4-HMLT as a secondary antioxidant. Albeit they were predicted to be efficient protectors against OS, the computationally-designed molecules have not been synthesized yet, so their AOC can be tested and the predictions might be confirmed or refuted.
Other challenges in this context are identifying the products yielded from the chemical AOC of phenolic MLT-related compounds (under physiological conditions), their relative abundance and their chemical fate. Further investigations are also desired regarding the possibility that they might be pro-oxidants under certain conditions, and their potential interactions with frequently used medical drugs.

Acknowledgments

A.P.-G. acknowledges the economic support of the Program of Cátedras—CONACYT from CONACyT—UAMI (2015–2025); and R.C.-A. acknowledges the financial support from CONACyT postdoctoral fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Davies, K.J. Oxidative stress: The paradox of aerobic life. Biochem. Soc. Symp. 1995, 61, 1–31. [Google Scholar] [CrossRef] [PubMed]
  2. Pham-Huy, L.A.; He, H.; Pham-Huy, C. Free radicals, antioxidants in disease and health. Int. J. Biomed. Sci. 2008, 4, 89–96. [Google Scholar] [PubMed]
  3. Pacher, P.; Beckman, J.S.; Liaudet, L. Nitric oxide and peroxynitrite in health and disease. Physiol. Rev. 2007, 87, 315–424. [Google Scholar] [CrossRef] [PubMed]
  4. Genestra, M. Oxyl radicals, redox-sensitive signalling cascades and antioxidants. Cell. Signal. 2007, 19, 1807–1819. [Google Scholar] [CrossRef] [PubMed]
  5. 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. [Google Scholar] [CrossRef] [PubMed]
  6. 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. [Google Scholar] [CrossRef] [PubMed]
  7. Droge, W. Free radicals in the physiological control of cell function. Physiol. Rev. 2002, 82, 47–95. [Google Scholar] [CrossRef] [PubMed]
  8. Young, I.S.; Woodside, J.V. Antioxidants in health and disease. J. Clin. Pathol. 2001, 54, 176–186. [Google Scholar] [PubMed]
  9. Halliwell, B. Biochemistry of oxidative stress. Biochem. Soc. Trans. 2007, 35, 1147–1150. [Google Scholar] [CrossRef] [PubMed]
  10. Willcox, J.K.; Ash, S.L.; Catignani, G.L. Antioxidants and prevention of chronic disease. Crit. Rev. Food Sci. Nutr. 2004, 44, 275–295. [Google Scholar] [CrossRef] [PubMed]
  11. Parthasarathy, S.; Santanam, N.; Ramachandran, S.; Meilhac, O. Oxidants and antioxidants in atherogenesis: An appraisal. J. Lipid Res. 1999, 40, 2143–2157. [Google Scholar] [PubMed]
  12. Lowe, F.J.; Cemeli, E. Biomarkers of oxidative stress and the relationship to cigarette smoking. Mini Rev. Org. Chem. 2011, 8, 377–386. [Google Scholar] [CrossRef]
  13. Valko, M.; Morris, H.; Cronin, M.T.D. Metals, toxicity and oxidative stress. Curr. Med. Chem. 2005, 12, 1161–1208. [Google Scholar] [CrossRef] [PubMed]
  14. Reiter, R.J.; Tan, D.X.; Sainz, R.M.; Mayo, J.C.; Lopez-Burillo, S. Melatonin: Reducing the toxicity and increasing the efficacy of drugs. J. Pharm. Pharmacol. 2002, 54, 1299–1321. [Google Scholar] [CrossRef] [PubMed]
  15. Reiter, R.J.; Manchester, L.C.; Tan, D.X. Neurotoxins: Free radical mechanisms and melatonin protection. Curr. Neuropharmacol. 2010, 8, 194–210. [Google Scholar] [CrossRef] [PubMed]
  16. Thanan, R.; Oikawa, S.; Hiraku, Y.; Ohnishi, S.; Ma, N.; Pinlaor, S.; Yongvanit, P.; Kawanishi, S.; Murata, M. Oxidative stress and its significant roles in neurodegenerative diseases and cancer. Int. J. Mol. Sci. 2014, 16, 193–217. [Google Scholar] [CrossRef] [PubMed]
  17. Filaire, E.; Dupuis, C.; Galvaing, G.; Aubreton, S.; Laurent, H.; Richard, R.; Filaire, M. Lung cancer: What are the links with oxidative stress, physical activity and nutrition. Lung Cancer 2013, 82, 383–389. [Google Scholar] [CrossRef] [PubMed]
  18. Paschos, A.; Pandya, R.; Duivenvoorden, W.C.M.; Pinthus, J.H. Oxidative stress in prostate cancer: Changing research concepts towards a novel paradigm for prevention and therapeutics. Prostate Cancer Prostatic Dis. 2013, 16, 217–225. [Google Scholar] [CrossRef] [PubMed]
  19. Siti, H.N.; Kamisah, Y.; Kamsiah, J. The role of oxidative stress, antioxidants and vascular inflammation in cardiovascular disease (a review). Vasc. Pharmacol. 2015, 71, 40–56. [Google Scholar] [CrossRef] [PubMed]
  20. Popolo, A.; Autore, G.; Pinto, A.; Marzocco, S. Oxidative stress in patients with cardiovascular disease and chronic renal failure. Free Radic. Res. 2013, 47, 346–356. [Google Scholar] [CrossRef] [PubMed]
  21. Miller, M.R.; Shaw, C.A.; Langrish, J.P. From particles to patients: Oxidative stress and the cardiovascular effects of air pollution. Future Cardiol. 2012, 8, 577–602. [Google Scholar] [CrossRef] [PubMed]
  22. Blesa, J.; Trigo-Damas, I.; Quiroga-Varela, A.; Jackson-Lewis, V.R. Oxidative stress and Parkinson’s disease. Front. Neuroanat. 2015, 9, 91. [Google Scholar] [CrossRef] [PubMed]
  23. Sharma, N.; Nehru, B. Characterization of the lipopolysaccharide induced model of Parkinson’s disease: Role of oxidative stress and neuroinflammation. Neurochem. Int. 2015, 87, 92–105. [Google Scholar] [CrossRef] [PubMed]
  24. Gaki, G.S.; Papavassiliou, A.G. Oxidative stress-induced signaling pathways implicated in the pathogenesis of Parkinson’s disease. Neuromol. Med. 2014, 16, 217–230. [Google Scholar] [CrossRef] [PubMed]
  25. Swomley, A.M.; Butterfield, D.A. Oxidative stress in Alzheimer disease and mild cognitive impairment: Evidence from human data provided by redox proteomics. Arch. Toxicol. 2015, 89, 1669–1680. [Google Scholar] [CrossRef] [PubMed]
  26. Mota, S.I.; Costa, R.O.; Ferreira, I.L.; Santana, I.; Caldeira, G.L.; Padovano, C.; Fonseca, A.C.; Baldeiras, I.; Cunha, C.; Letra, L.; et al. Oxidative stress involving changes in Nrf2 and ER stress in early stages of Alzheimer’s disease. Biochim. Biophys. Acta 2015, 1852, 1428–1441. [Google Scholar] [CrossRef] [PubMed]
  27. Meraz-Ríos, M.A.; Franco-Bocanegra, D.; Toral Rios, D.; Campos-Peña, V. Early onset Alzheimer’s disease and oxidative stress. Oxid. Med. Cell. Longev. 2014, 2014, 375968. [Google Scholar] [CrossRef] [PubMed]
  28. Tousoulis, D.; Psaltopoulou, T.; Androulakis, E.; Papageorgiou, N.; Papaioannou, S.; Oikonomou, E.; Synetos, A.; Stefanadis, C. Oxidative stress and early atherosclerosis: Novel antioxidant treatment. Cardiovasc. Drugs Ther. 2015, 29, 75–88. [Google Scholar] [CrossRef] [PubMed]
  29. Li, H.; Horke, S.; Förstermann, U. Vascular oxidative stress, nitric oxide and atherosclerosis. Atherosclerosis 2014, 237, 208–219. [Google Scholar] [CrossRef] [PubMed]
  30. Peluso, I.; Morabito, G.; Urban, L.; Ioannone, F.; Serafini, M. Oxidative stress in atherosclerosis development: The central role of LDL and oxidative burst. Endocr. Metab. Immune Disord. 2012, 12, 351–360. [Google Scholar] [CrossRef]
  31. Rosales-Corral, S.; Tan, D.X.; Manchester, L.; Reiter, R.J. Diabetes and Alzheimer disease, two overlapping pathologies with the same background: Oxidative stress. Oxid. Med. Cell. Longev. 2015, 2015, 985845. [Google Scholar] [CrossRef] [PubMed]
  32. Zephy, D.; Ahmad, J. Type 2 diabetes mellitus: Role of melatonin and oxidative stress. Diabetes Metab. Syndr. 2015, 9, 127–131. [Google Scholar] [CrossRef] [PubMed]
  33. Maiese, K. New insights for oxidative stress and diabetes mellitus. Oxid. Med. Cell. Longev. 2015, 2015, 875961. [Google Scholar] [CrossRef] [PubMed]
  34. Ramis, M.R.; Esteban, S.; Miralles, A.; Tan, D.X.; Reiter, R.J. Protective effects of melatonin and mitochondria-targeted antioxidants against oxidative stress: A review. Curr. Med. Chem. 2015, 22, 2690–2711. [Google Scholar] [CrossRef] [PubMed]
  35. Galano, A.; Tan, D.X.; Reiter, R.J. Melatonin as a natural ally against oxidative stress: A physicochemical examination. J. Pineal Res. 2011, 51, 1–16. [Google Scholar] [CrossRef] [PubMed]
  36. Hardeland, R. Antioxidative protection by melatonin: Multiplicity of mechanisms from radical detoxification to radical avoidance. Endocrine 2005, 27, 119–130. [Google Scholar] [CrossRef]
  37. Jahnke, G.; Marr, M.; Myers, C.; Wilson, R.; Travlos, G.; Price, C. Maternal and developmental toxicity evaluation of melatonin administered orally to pregnant Sprague-Dawley rats. Toxicol. Sci. 1999, 50, 271–279. [Google Scholar] [CrossRef] [PubMed]
  38. Ceraulo, L.; Ferrugia, M.; Tesoriere, L.; Segreto, S.; Livrea, M.A.; Turco Liveri, V. Interactions of melatonin with membrane models: Portioning of melatonin in AOT and lecithin reversed micelles. J. Pineal Res. 1999, 26, 108–112. [Google Scholar] [CrossRef] [PubMed]
  39. Bonnefont-Rousselot, D.; Collin, F. Melatonin: Action as antioxidant and potential applications in human disease and aging. Toxicology 2010, 278, 55–67. [Google Scholar] [CrossRef] [PubMed]
  40. Gurer-Orhan, H.; Suzen, S. Melatonin, its metabolites and its synthetic analogs as multi-faceted compounds: Antioxidant, prooxidant and inhibitor of bioactivation reactions. Curr. Med. Chem. 2015, 22, 490–499. [Google Scholar] [CrossRef] [PubMed]
  41. 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. [Google Scholar] [CrossRef] [PubMed]
  42. Reiter, R.J.; Tan, D.X.; Jou, M.J.; Korkmaz, A.; Manchester, L.C.; Paredes, S.D. Biogenic amines in the reduction of oxidative stress: Melatonin and its metabolites. Neuro Endocrinol. Lett. 2008, 29, 391–398. [Google Scholar] [PubMed]
  43. Álvarez-Diduk, R.; Galano, A.; Tan, D.X.; Reiter, R.J. N-Acetylserotonin and 6-hydroxymelatonin against oxidative stress: Implications for the overall protection exerted by melatonin. J. Phys. Chem. B 2015, 119, 8535–8543. [Google Scholar] [CrossRef] [PubMed]
  44. Tan, D.X.; Hardeland, R.; Manchester, L.C.; Galano, A.; Reiter, R.J. Cyclic-3-hydroxymelatonin (C3HOM), a potent antioxidant, scavenges free radicals and suppresses oxidative reactions. Curr. Med. Chem. 2014, 21, 1557–1565. [Google Scholar] [CrossRef] [PubMed]
  45. Galano, A.; Tan, D.X.; Reiter, R.J. Cyclic 3-hydroxymelatonin, a key metabolite enhancing the peroxyl radical scavenging activity of melatonin. RSC Adv. 2014, 4, 5220–5227. [Google Scholar] [CrossRef]
  46. Martins, N.; Barros, L.; Ferreira, I.C.F.R. In vivo antioxidant activity of phenolic compounds: Facts and gaps. Trends Food Sci. Technol. 2016, 48, 1–12. [Google Scholar] [CrossRef] [Green Version]
  47. Achat, S.; Rakotomanomana, N.; Madani, K.; Dangles, O. Antioxidant activity of olive phenols and other dietary phenols in model gastric conditions: Scavenging of the free radical DPPH and inhibition of the haem-induced peroxidation of linoleic acid. Food Chem. 2016, 213, 135–142. [Google Scholar] [CrossRef] [PubMed]
  48. Shahidi, F.; Ambigaipalan, P. Phenolics and polyphenolics in foods, beverages and spices: Antioxidant activity and health effects—A review. J. Funct. Foods 2015, 18, 820–897. [Google Scholar] [CrossRef]
  49. 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. 2015, 59, 231–262. [Google Scholar]
  50. Araújo, M.; Pimentel, F.B.; Alves, R.C.; Oliveira, M.B.P.P. Phenolic compounds from olive mill wastes: Health effects, analytical approach and application as food antioxidants. Trends Food Sci. Technol. 2015, 45, 200–211. [Google Scholar] [CrossRef]
  51. Liu, L.; Jin, C.; Zhang, Y. Lipophilic phenolic compounds (Lipo-PCs): Emerging antioxidants applied in lipid systems. RSC Adv. 2014, 4, 2879–2891. [Google Scholar] [CrossRef]
  52. Ho, C.T.; Wang, M. Dietary phenolics as reactive carbonyl scavengers: Potential impact on human health and mechanism of action. J. Tradit. Complement. Med. 2013, 3, 139–141. [Google Scholar] [CrossRef] [PubMed]
  53. Galano, A. Computational-aided design of melatonin analogues with outstanding multifunctional antioxidant capacity. RSC Adv. 2016, 6, 22951–22963. [Google Scholar] [CrossRef]
  54. Gerschman, R.; Gilbert, D.L.; Nye, S.W.; Dwyer, P.; Fenn, W.O. Oxygen poisoning and X-irradiation: A mechanism in common. Science 1954, 119, 623–626. [Google Scholar] [CrossRef] [PubMed]
  55. Kaminskyy, V.O.; Zhivotovsky, B. Free radicals in cross talk between autophagy and apoptosis. Antioxid. Redox Signal. 2014, 21, 86–102. [Google Scholar] [CrossRef] [PubMed]
  56. Yin, H.; Zhu, M. Free radical oxidation of cardiolipin: Chemical mechanisms, detection and implication in apoptosis, mitochondrial dysfunction and human diseases. Free Radic. Res. 2012, 46, 959–974. [Google Scholar] [CrossRef] [PubMed]
  57. Sayre, L.M.; Perry, G.; Smith, M.A. Oxidative stress and neurotoxicity. Chem. Res. Toxicol. 2008, 21, 172–188. [Google Scholar] [CrossRef] [PubMed]
  58. Morris, G.; Walder, K.; Puri, B.K.; Berk, M.; Maes, M. The Deleterious Effects of Oxidative and Nitrosative Stress on Palmitoylation, Membrane Lipid Rafts and Lipid-Based Cellular Signalling: New Drug Targets in Neuroimmune Disorders. Mol. Neurobiol. 2016, 53, 4638–4658. [Google Scholar] [CrossRef] [PubMed]
  59. Niedzielska, E.; Smaga, I.; Gawlik, M.; Moniczewski, A.; Stankowicz, P.; Pera, J.; Filip, M. Oxidative Stress in Neurodegenerative Diseases. Mol. Neurobiol. 2016, 53, 4094–4125. [Google Scholar] [CrossRef] [PubMed]
  60. Daulatzai, M.A. Fundamental role of pan-inflammation and oxidative-nitrosative pathways in neuropathogenesis of Alzheimer’s disease. Am. J. Neurodegener. Dis. 2016, 5, 1–28. [Google Scholar] [PubMed]
  61. Ljubisavljevic, S. Oxidative Stress and Neurobiology of Demyelination. Mol. Neurobiol. 2016, 53, 744–758. [Google Scholar] [CrossRef] [PubMed]
  62. Cobb, C.A.; Cole, M.P. Oxidative and nitrative stress in neurodegeneration. Neurobiol. Dis. 2015, 84, 4–21. [Google Scholar] [CrossRef] [PubMed]
  63. Bhat, A.H.; Dar, K.B.; Anees, S.; Zargar, M.A.; Masood, A.; Sofi, M.A.; Ganie, S.A. Oxidative stress, mitochondrial dysfunction and neurodegenerative diseases; a mechanistic insight. Biomed. Pharmacother. 2015, 74, 101–110. [Google Scholar] [CrossRef] [PubMed]
  64. Fischer, R.; Maier, O. Interrelation of oxidative stress and inflammation in neurodegenerative disease: Role of TNF. Oxid. Med. Cell. Longev. 2015, 2015, 610813. [Google Scholar] [CrossRef] [PubMed]
  65. Hecht, F.; Pessoa, C.F.; Gentile, L.B.; Rosenthal, D.; Carvalho, D.P.; Fortunato, R.S. The role of oxidative stress on breast cancer development and therapy. Tumor Biol. 2016, 37, 4281–4291. [Google Scholar] [CrossRef] [PubMed]
  66. Zuo, T.; Zhu, M.; Xu, W. Roles of oxidative stress in polycystic ovary syndrome and cancers. Oxid. Med. Cell. Longev. 2016, 2016, 8589318. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, Z.; Li, S.; Cao, Y.; Tian, X.; Zeng, R.; Liao, D.F.; Cao, D. Oxidative stress and carbonyl lesions in ulcerative colitis and associated colorectal cancer. Oxid. Med. Cell. Longev. 2016, 2016, 9875298. [Google Scholar] [CrossRef] [PubMed]
  68. Dizdaroglu, M. Oxidatively induced DNA damage and its repair in cancer. Mutat. Res. Rev. Mutat. Res. 2015, 763, 212–245. [Google Scholar] [CrossRef] [PubMed]
  69. Choudhari, S.K.; Chaudhary, M.; Gadbail, A.R.; Sharma, A.; Tekade, S. Oxidative and antioxidative mechanisms in oral cancer and precancer: A review. Oral Oncol. 2014, 50, 10–18. [Google Scholar] [CrossRef] [PubMed]
  70. Kayama, Y.; Raaz, U.; Jagger, A.; Adam, M.; Schellinger, I.N.; Sakamoto, M.; Suzuki, H.; Toyama, K.; Spin, J.M.; Tsao, P.S. Diabetic cardiovascular disease induced by oxidative stress. Int. J. Mol. Sci. 2015, 16, 25234–25263. [Google Scholar] [CrossRef] [PubMed]
  71. Roul, D.; Recchia, F.A. Metabolic Alterations Induce Oxidative Stress in Diabetic and Failing Hearts: Different Pathways, Same Outcome. Antioxid. Redox Signal. 2015, 22, 1502–1514. [Google Scholar] [CrossRef] [PubMed]
  72. Marseglia, L.; D’Angelo, G.; Manti, S.; Arrigo, T.; Barberi, I.; Reiter, R.J.; Gitto, E. Oxidative stress-mediated aging during the fetal and perinatal periods. Oxid. Med. Cell. Longev. 2014, 2014, 358375. [Google Scholar] [CrossRef] [PubMed]
  73. Wells, P.G.; Bhatia, S.; Drake, D.M.; Miller-Pinsler, L. Fetal oxidative stress mechanisms of neurodevelopmental deficits and exacerbation by ethanol and methamphetamine. Birth Defects Res. C Embryo Today 2016, 108, 108–130. [Google Scholar] [CrossRef] [PubMed]
  74. Hansson, S.R.; Nääv, A.; Erlandsson, L. Oxidative stress in preeclampsia and the role of free fetal hemoglobin. Front. Physiol. 2015, 6, 516. [Google Scholar] [CrossRef] [PubMed]
  75. Perrone, S.; Tataranno, M.L.; Stazzoni, G.; Buonocore, G. Biomarkers of oxidative stress in fetal and neonatal diseases. J. Matern. Fetal Neonatal Med. 2012, 25, 2575–2578. [Google Scholar] [CrossRef] [PubMed]
  76. Correia-Costa, L.; Sousa, T.; Morato, M.; Cosme, D.; Afonso, J.; Areias, J.C.; Schaefer, F.; Guerra, A.; Afonso, A.C.; Azevedo, A.; et al. Oxidative stress and nitric oxide are increased in obese children and correlate with cardiometabolic risk and renal function. Br. J. Nutr. 2016, 116, 805–815. [Google Scholar] [CrossRef] [PubMed]
  77. Di Pietro, M.; Filardo, S.; de Santis, F.; Mastromarino, P.; Sessa, R. Chlamydia pneumoniae and oxidative stress in cardiovascular disease: State of the art and prevention strategies. Int. J. Mol. Sci. 2015, 16, 724–735. [Google Scholar] [CrossRef] [PubMed]
  78. Wu, J.; Xia, S.; Kalionis, B.; Wan, W.; Sun, T. The Role of Oxidative Stress and Inflammation in Cardiovascular Aging. Biomed. Res. Int. 2014, 2014, 615312. [Google Scholar] [CrossRef] [PubMed]
  79. Zhang, P.Y.; Xu, X.; Li, X.C. Cardiovascular diseases: Oxidative damage and antioxidant protection. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 3091–3096. [Google Scholar] [PubMed]
  80. Ferrari, R.S.; Andrade, C.F. Oxidative Stress and Lung Ischemia-Reperfusion Injury. Oxid. Med. Cell. Longev. 2015, 2015, 590987. [Google Scholar] [CrossRef] [PubMed]
  81. Cheresh, P.; Kim, S.J.; Tulasiram, S.; Kamp, D.W. Oxidative stress and pulmonary fibrosis. Biochim. Biophys. Acta 2013, 1832, 1028–1040. [Google Scholar] [CrossRef] [PubMed]
  82. 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. [Google Scholar] [CrossRef] [PubMed]
  83. Ratliff, B.B.; Abdulmahdi, W.; Pawar, R.; Wolin, M.S. Oxidant mechanisms in renal injury and disease. Antioxid. Redox Signal. 2016, 25, 119–146. [Google Scholar] [CrossRef] [PubMed]
  84. Tamay-Cach, F.; Quintana-Pérez, J.C.; Trujillo-Ferrara, J.G.; Cuevas-Hernández, R.I.; Del Valle-Mondragón, L.; García-Trejo, E.M.; Arellano-Mendoza, M.G. A review of the impact of oxidative stress and some antioxidant therapies on renal damage. Ren. Fail. 2016, 38, 171–175. [Google Scholar] [CrossRef] [PubMed]
  85. Nita, M.; Grzybowski, A. The Role of the Reactive Oxygen Species and Oxidative Stress in the Pathomechanism of the Age-Related Ocular Diseases and Other Pathologies of the Anterior and Posterior Eye Segments in Adults. Oxid. Med. Cell. Longev. 2016, 2016, 3164734. [Google Scholar] [CrossRef] [PubMed]
  86. Kruk, J.; Kubasik-Kładna, K.; Aboul-Enein, H.Y. The role oxidative stress in the pathogenesis of eye diseases: Current status and a dual role of physical activity. Mini-Rev. Med. Chem. 2016, 16, 241–257. [Google Scholar] [CrossRef]
  87. Saccà, S.C.; Izzotti, A. Oxidative stress and glaucoma: Injury in the anterior segment of the eye. Prog. Brain Res. 2008, 173, 385–407. [Google Scholar] [PubMed]
  88. Griffiths, H.R.; Dunston, C.R.; Bennett, S.J.; Grant, M.M.; Phillips, D.C.; Kitas, G.D. Free radicals and redox signalling in T-cells during chronic inflammation and ageing. Biochem. Soc. Trans. 2011, 39, 1273–1278. [Google Scholar] [CrossRef] [PubMed]
  89. Schetter, A.J.; Heegaard, N.H.H.; Harris, C.C. Inflammation and cancer: Interweaving microRNA, free radical, cytokine and p53 pathways. Carcinogenesis 2009, 31, 37–49. [Google Scholar] [CrossRef] [PubMed]
  90. Giannapas, M.; Karnis, L.; Dailianis, S. Generation of free radicals in haemocytes of mussels after exposure to low molecular weight PAH components: Immune activation, oxidative and genotoxic effects. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2012, 155, 182–189. [Google Scholar] [CrossRef] [PubMed]
  91. Lantow, M.; Schuderer, J.; Hartwig, C.; Simkó, M. Free radical release and HSP70 expression in two human immune-relevant cell lines after exposure to 1800 MHz radiofrequency radiation. Radiat. Res. 2006, 165, 88–94. [Google Scholar] [CrossRef] [PubMed]
  92. 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. 2015, 125, 555–565. [Google Scholar] [CrossRef] [PubMed]
  93. Niatsetskaya, Z.V.; Sosunov, S.A.; Matsiukevich, D.; Utkina-Sosunova, I.V.; Ratner, V.I.; Starkov, A.A.; Ten, V.S. The oxygen free radicals originating from mitochondrial complex I contribute to oxidative brain injury following hypoxia-ischemia in neonatal mice. J. Neurosci. 2012, 32, 3235–3244. [Google Scholar] [CrossRef] [PubMed]
  94. Maeda, H. The link between infection and cancer: Tumor vasculature, free radicals, and drug delivery to tumors via the EPR effect. Cancer Sci. 2013, 104, 779–789. [Google Scholar] [CrossRef] [PubMed]
  95. Guha, M.; Maity, P.; Choubey, V.; Mitra, K.; Reiter, R.J.; Bandyopadhyay, U. Melatonin inhibits free radical-mediated mitochondrial-dependent hepatocyte apoptosis and liver damage induced during malarial infection. J. Pineal Res. 2007, 43, 372–381. [Google Scholar] [CrossRef] [PubMed]
  96. De Luca, C.; Deeva, I.; Mariani, S.; Maiani, G.; Stancato, A.; Korkina, L. Monitoring antioxidant defenses and free radical production in space-flight, aviation and railway engine operators, for the prevention and treatment of oxidative stress, immunological impairment, and pre-mature cell aging. Toxicol. Ind. Health 2009, 25, 259–267. [Google Scholar] [CrossRef] [PubMed]
  97. Morales-Alamo, D.; Calbet, J.A.L. Free radicals and sprint exercise in humans. Free Radic. Res. 2014, 48, 30–42. [Google Scholar] [CrossRef] [PubMed]
  98. Benameur, L.; Charif, N.; Li, Y.; Stoltz, J.F.; de Isla, N. Toward an understanding of mechanism of aging-induced oxidative stress in human mesenchymal stem cells. Bio-Med. Mater. Eng. 2015, 25, S41–S46. [Google Scholar]
  99. Shiraiwa, M.; Selzle, K.; Pöschl, U. Hazardous components and health effects of atmospheric aerosol particles: Reactive oxygen species, soot, polycyclic aromatic compounds and allergenic proteins. Free Radic. Res. 2012, 46, 927–939. [Google Scholar] [CrossRef] [PubMed]
  100. Valencia-Islas, N.; Zambrano, A.; Rojas, J.L. Ozone reactivity and free radical scavenging behavior of phenolic secondary metabolites in lichens exposed to chronic oxidant air pollution from Mexico City. J. Chem. Ecol. 2007, 33, 1619–1634. [Google Scholar] [CrossRef] [PubMed]
  101. Kamiński, P.; Kurhalyuk, N.; Szady-Grad, M. Heavy metal-induced oxidative stress and changes in physiological process of free radicals in the blood of white stork (Ciconia ciconia) chicks in polluted areas. Pol. J. Environ. Stud. 2007, 16, 555–562. [Google Scholar]
  102. Vejerano, E.; Lomnicki, S.; Dellinger, B. Lifetime of combustion-generated environmentally persistent free radicals on Zn(II)O and other transition metal oxides. J. Environ. Monit. 2012, 14, 2803–2806. [Google Scholar] [CrossRef] [PubMed]
  103. Robinson, E.A.; Johnson, J.D. Methods for analysis of free radicals in cigarette smoke. Mini Rev. Org. Chem. 2011, 8, 401–411. [Google Scholar] [CrossRef]
  104. Wang, Y.; Zhang, Y.; Wu, D.; Liu, B.; Xie, W. Electron paramagnetic resonance spin-trapping study of free radicals in gas phase of mainstream cigarette smoke using two spin traps. Tobacco Sci. Technol. 2014, 47, 47–53. [Google Scholar]
  105. Michail, K.; Baghdasarian, A.; Narwaley, M.; Aljuhani, N.; Siraki, A.G. Scavenging of free-radical metabolites of aniline xenobiotics and drugs by amino acid derivatives: Toxicological implications of radical-transfer reactions. Chem. Res. Toxicol. 2013, 26, 1872–1883. [Google Scholar] [CrossRef] [PubMed]
  106. Aleryani, S.L.; Aleryani, R.A.; Al-Akwa, A.A. Khat a drug of abuse: Roles of free radicals and antioxidants. Drug Test. Anal. 2011, 3, 548–551. [Google Scholar] [CrossRef] [PubMed]
  107. Narwaley, M.; Michail, K.; Arvadia, P.; Siraki, A.G. Drug-induced protein free radical formation is attenuated by unsaturated fatty acids by scavenging drug-derived phenyl radical metabolites. Chem. Res. Toxicol. 2011, 24, 1031–1039. [Google Scholar] [CrossRef] [PubMed]
  108. Karadayian, A.G.; Bustamante, J.; Czerniczyniec, A.; Lombardi, P.; Cutrera, R.A.; Lores-Arnaiz, S. Alcohol hangover induces mitochondrial dysfunction and free radical production in mouse cerebellum. Neuroscience 2015, 304, 47–59. [Google Scholar] [CrossRef] [PubMed]
  109. Albano, E. Alcohol, oxidative stress and free radical damage. Proc. Nutr. Soc. 2006, 65, 278–290. [Google Scholar] [CrossRef] [PubMed]
  110. Spitz, D.R.; Hauer-Jensen, M. Ionizing radiation-induced responses: Where free radical chemistry meets redox biology and medicine. Antioxid. Redox Signal. 2014, 20, 1407–1409. [Google Scholar] [CrossRef] [PubMed]
  111. Rancan, F.; Nazemi, B.; Rautenberg, S.; Ryll, M.; Hadam, S.; Gao, Q.; Hackbarth, S.; Haag, S.F.; Graf, C.; Rühl, E.; et al. Ultraviolet radiation and nanoparticle induced intracellular free radicals generation measured in human keratinocytes by electron paramagnetic resonance spectroscopy. Skin Res. Technol. 2014, 20, 182–193. [Google Scholar] [CrossRef] [PubMed]
  112. Burlaka, A.; Tsybulin, O.; Sidorik, E.; Lukin, S.; Polishuk, V.; Tsehmistrenko, S.; Yakymenko, I. Overproduction of free radical species in embryonal cells exposed to low intensity radiofrequency radiation. Exp. Oncol. 2013, 35, 219–225. [Google Scholar] [PubMed]
  113. Milnerowicz, H.; Ściskalska, M.; Dul, M. Molecular mechanisms of the impact of smoke-oxidants. Exp. Toxicol. Pathol. 2015, 67, 377–382. [Google Scholar] [CrossRef] [PubMed]
  114. Buxton, G.V.; Greenstock, C.L.; Helman, W.P.; Ross, A.B. Critical Review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (OH/O in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17, 513–886. [Google Scholar] [CrossRef]
  115. Rose, R.C.; Bode, A.M. Biology of free radical scavengers: An evaluation of ascorbate. FASEB J. 1993, 7, 1135–1142. [Google Scholar] [PubMed]
  116. Reiter, R.J.; Tan, D.X.; Herman, T.S.; Thomas, C.R., Jr. Melatonin as a radioprotective agent: A review. Int. J. Radiat. Oncol. Biol. Phys. 2004, 59, 639–653. [Google Scholar]
  117. Candeias, L.P.; Steenken, S. Reaction of HO with guanine derivatives in aqueous solution: Formation of two different redox-active OH-adduct radicals and their unimolecular transformation reactions. Properties of G(-H). Chem. Eur. J. 2000, 6, 475–484. [Google Scholar] [CrossRef]
  118. Chatgilialoglu, C.; D’Angelantonio, M.; Guerra, M.; Kaloudis, P.; Mulazzani, Q.G. A reevaluation of the ambident reactivity of the guanine moiety towards hydroxyl radicals. Angew. Chem. Int. Ed. Engl. 2009, 48, 2214–2217. [Google Scholar] [CrossRef] [PubMed]
  119. Galano, A.; Alvarez-Idaboy, J.R. Guanosine + OH radical reaction in aqueous solution: A reinterpretation of the UV-vis data based on thermodynamic and kinetic calculations. Org. Lett. 2009, 11, 5114–5117. [Google Scholar] [CrossRef] [PubMed]
  120. León-Carmona, J.R.; Galano, A. Is caffeine a good scavenger of oxygenated free radicals? J. Phys. Chem. B 2011, 115, 4538–4546. [Google Scholar] [CrossRef] [PubMed]
  121. Galano, A. Relative antioxidant efficiency of a large series of carotenoids in terms of one electron transfer reactions. J. Phys. Chem. B 2007, 111, 12898–12908. [Google Scholar] [CrossRef] [PubMed]
  122. Martinez, A.; Vargas, R.; Galano, A. What is important to prevent oxidative stress? A theoretical study on electron-transfer reactions between carotenoids and free radicals. J. Phys. Chem. B 2009, 113, 12113–12120. [Google Scholar] [CrossRef] [PubMed]
  123. Martínez, A.; Vargas, R.; Galano, A. Theoretical study on the chemical fate of adducts formed through free radical addition reactions to carotenoids. Theor. Chem. Acc. 2010, 127, 595–603. [Google Scholar] [CrossRef]
  124. Martínez, A.; Galano, A. Free radical scavenging activity of ultrashort single-walled carbon nanotubes with different structures through electron transfer reactions. J. Phys. Chem. C 2010, 114, 8184–8191. [Google Scholar] [CrossRef]
  125. Marnett, L.J. Peroxyl free radicals: Potential mediators of tumor initiation and promotion. Carcinogenesis 1987, 8, 1365–1373. [Google Scholar] [CrossRef] [PubMed]
  126. Terpinc, P.; Abramovič, H. A kinetic approach for evaluation of the antioxidant activity of selected phenolic acids. Food Chem. 2010, 121, 366–371. [Google Scholar] [CrossRef]
  127. Sies, H. Oxidative stress: Oxidants and antioxidants. Exp. Physiol. 1997, 82, 291–295. [Google Scholar] [CrossRef] [PubMed]
  128. De Grey, A.D.N.J. HO2: The forgotten radical. DNA Cell Biol. 2002, 21, 251–257. [Google Scholar] [CrossRef] [PubMed]
  129. Bielski, B.H.J.; Arudi, R.L.; Sutherland, M.W. A study of the reactivity of HO2/O2—With unsaturated fatty acids. J. Biol. Chem. 1983, 258, 4759–4761. [Google Scholar] [PubMed]
  130. Aikens, J.; Dix, T.A. Perhydroxyl radical (HOO) initiated lipid peroxidation: The role of fatty acid hydroperoxides. J. Biol. Chem. 1991, 266, 15091–15098. [Google Scholar] [PubMed]
  131. Galano, A. On the direct scavenging activity of melatonin towards hydroxyl and a series of peroxyl radicals. Phys. Chem. Chem. Phys. 2011, 13, 7178–7188. [Google Scholar] [CrossRef] [PubMed]
  132. Neta, P.; Huie, R.E.; Mosseri, S.; Shastri, L.V.; Mittal, J.P.; Maruthamuthu, P.; Steenken, S. Rate constants for reduction of substituted methylperoxyl radicals by ascorbate ions and N,N,N′,N′,-Tetramethyl-p-phenylenediamine. J. Phys. Chem. 1989, 93, 4099–4104. [Google Scholar] [CrossRef]
  133. Brunelli, L.; Crow, J.P.; Beckman, J.S. The comparative toxicity of nitric oxide and peroxynitrite to Escherichia coli. Arch. Biochem. Biophys. 1995, 316, 327–334. [Google Scholar] [CrossRef] [PubMed]
  134. Squadrito, G.L.; Pryor, W.A. Oxidative chemistry of nitric oxide: The roles of superoxide, peroxynitrite, and carbon dioxide. Free Radic. Biol. Med. 1998, 25, 392–403. [Google Scholar] [CrossRef]
  135. Radi, R.; Peluffo, G.; Alvarez, M.N.; Naviliat, M.; Cayota, A. Unraveling peroxynitrite formation in biological systems. Free Radic. Biol. Med. 2001, 30, 463–488. [Google Scholar] [CrossRef]
  136. Douki, T.; Cadet, J. Peroxynitrite mediated oxidation of purine bases of nucleosides and isolated DNA. Free Radic. Res. 1996, 24, 369–380. [Google Scholar] [CrossRef] [PubMed]
  137. Wiseman, H.; Halliwell, B. Damage to DNA by reactive oxygen and nitrogen species: Role in inflammatory disease and progression to cancer. Biochem. J. 1996, 313, 17–29. [Google Scholar] [CrossRef] [PubMed]
  138. Koppal, T.; Drake, J.; Yatin, S.; Jordan, B.; Varadarajan, S.; Bettenhausen, L.; Butterfield, D.A. Peroxynitrite-induced alterations in synaptosomal membrane proteins: Insight into oxidative stress in Alzheimer’s disease. J. Neurochem. 1999, 72, 310–317. [Google Scholar] [CrossRef] [PubMed]
  139. Prütz, W.A.; Mönig, H.; Butler, J.; Land, E.J. Reactions of nitrogen dioxide in aqueous model systems: Oxidation of tyrosine units in peptides and proteins. Arch. Biochem. Biophys. 1985, 243, 125–134. [Google Scholar] [CrossRef]
  140. Abedinzadeh, Z. Sulfur-centered reactive intermediates derived from the oxidation of sulfur compounds of biological interest. Can. J. Physiol. Pharmacol. 2001, 79, 166–170. [Google Scholar] [CrossRef] [PubMed]
  141. Giles, G.I.; Tasker, K.M.; Jacob, C. Hypothesis: The role of reactive sulfur species in oxidative stress. Free Radic. Biol. Med. 2001, 31, 1279–1283. [Google Scholar] [CrossRef]
  142. Giles, G.I.; Tasker, K.M.; Jacob, C. Oxidation of biological thiols by highly reactive disulfide-S-oxides. Gen. Physiol. Biophys. 2002, 21, 65–72. [Google Scholar] [PubMed]
  143. Mishanina, T.V.; Libiad, M.; Banerjee, R. Biogenesis of reactive sulfur species for signaling by hydrogen sulfide oxidation pathways. Nat. Chem. Biol. 2015, 11, 457–464. [Google Scholar] [CrossRef] [PubMed]
  144. Rao, A.V.; Agarwal, S. Role of lycopene as antioxidant carotenoid in the prevention of chronic diseases: A review. Nutr. Res. 1999, 19, 305–323. [Google Scholar] [CrossRef]
  145. Sita, G.; Hrelia, P.; Tarozzi, A.; Morroni, F. Isothiocyanates are promising compounds against oxidative stress, neuroinflammation and cell death that may benefit neurodegeneration in Parkinson’s disease. Int. J. Mol. Sci. 2016, 17, 1454. [Google Scholar] [CrossRef] [PubMed]
  146. Skibska, B.; Goraca, A. The protective effect of lipoic acid on selected cardiovascular diseases caused by age-related oxidative stress. Oxid. Med. Cell. Longev. 2015, 2015, 313021. [Google Scholar] [CrossRef] [PubMed]
  147. Lançon, A.; Frazzi, R.; Latruffe, N. Anti-oxidant, anti-inflammatory and anti-angiogenic properties of resveratrol in ocular diseases. Molecules 2016, 21, 304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Zhang, H.; Tsao, R. Dietary polyphenols, oxidative stress and antioxidant and anti-inflammatory effects. Curr. Opin. Food Sci. 2016, 8, 33–42. [Google Scholar] [CrossRef]
  149. Niki, E. Antioxidant capacity of foods for scavenging reactive oxidants and inhibition of plasma lipid oxidation induced by multiple oxidants. Food Funct. 2016, 7, 2156–2168. [Google Scholar] [CrossRef] [PubMed]
  150. Amir Aslani, B.; Ghobadi, S. Studies on oxidants and antioxidants with a brief glance at their relevance to the immune system. Life Sci. 2016, 146, 163–173. [Google Scholar] [CrossRef] [PubMed]
  151. Kovacic, P.; Somanathan, R.; Abadjian, M.C.Z. Natural monophenols as therapeutics, antioxidants and toxins; electron transfer, radicals and oxidative stress. Nat. Prod. J. 2015, 5, 142–151. [Google Scholar] [CrossRef]
  152. Chaiyasit, W.; Elias, R.J.; McClements, D.J.; Decker, E.A. Role of physical structures in bulk oils on lipid oxidation. Crit. Rev. Food Sci. Nutr. 2007, 47, 299–317. [Google Scholar] [CrossRef] [PubMed]
  153. Litwinienko, G.; Ingold, K.U. Abnormal solvent effects on hydrogen atom abstractions. 1. The reactions of phenols with 2,2-diphenyl-1-picrylhydrazyl (dpph) in alcohols. J. Org. Chem. 2003, 68, 3433–3438. [Google Scholar] [CrossRef] [PubMed]
  154. Weinstein, J.; Bielski, B.H.J. Kinetics of the interaction of HO2 and O2—Radicals with hydrogen peroxide. The Haber-Weiss reaction. J. Am. Chem. Soc. 1979, 101, 58–62. [Google Scholar] [CrossRef]
  155. Rodriguez-Rodriguez, C.; Sanchez de Groot, N.; Rimola, A.; Alvarez-Larena, A.; Lloveras, V.; Vidal-Gancedo, J.; Ventura, S.; Vendrell, J.; Sodupe, M.; Gonzalez-Duarte, P. Design, selection, and characterization of thioflavin-based intercalation compounds with metal chelating properties for application in alzheimer’s disease. J. Am. Chem. Soc. 2009, 131, 1436–1451. [Google Scholar] [CrossRef] [PubMed]
  156. Francisco-Marquez, M.; Aguilar-Fernández, M.; Galano, A. Anthranilic acid as a secondary antioxidant: Implications to the inhibition of OH production and the associated oxidative stress. Comput. Theor. Chem. 2016, 1077, 18–24. [Google Scholar] [CrossRef]
  157. Miche, H.; Brumas, V.; Berthon, G. Copper(II) interactions with nonsteroidal antiinflammatory agents. II. Anthranilic acid as a potential OH-inactivating ligand. J. Inorg. Biochem. 1997, 68, 27–38. [Google Scholar] [CrossRef]
  158. Gaubert, S.; Bouchaut, M.; Brumas, V.; Berthon, G. Copper-ligand interactions and physiological free radical processes. Part 3. Influence of histidine, salicylic acid and anthranilic acid on copper-driven Fenton chemistry in vitro. Free Radic. Res. 2000, 32, 451–461. [Google Scholar] [CrossRef] [PubMed]
  159. Berthon, G. Is copper pro- or anti-inflammatory? A reconciling view and a novel approach for the use of copper in the control of inflammation. Agents Actions 1993, 39, 210–217. [Google Scholar] [CrossRef] [PubMed]
  160. Terao, J.; Minami, Y.; Bando, N. Singlet molecular oxygen-quenching activity of carotenoids: Relevance to protection of the skin from photoaging. J. Clin. Biochem. Nutr. 2011, 48, 57–62. [Google Scholar] [CrossRef] [PubMed]
  161. Shahidi, F.; Zhong, Y. Measurement of antioxidant activity. J. Funct. Foods 2015, 18, 757–781. [Google Scholar] [CrossRef]
  162. 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] [PubMed]
  163. Frankel, E.N.; Meyer, A.S. The problems of using one-dimensional methods to evaluate multifunctional food and biological antioxidants. J. Sci. Food Agric. 2000, 80, 1925–1941. [Google Scholar] [CrossRef]
  164. Antolovich, M.; Prenzler, P.D.; Patsalides, E.; McDonald, S.; Robards, K. Methods for testing antioxidant activity. Analyst 2002, 127, 183–198. [Google Scholar] [CrossRef] [PubMed]
  165. Huang, D.; Boxin, O.U.; Prior, R.L. The chemistry behind antioxidant capacity assays. J. Agric. Food Chem. 2005, 53, 1841–1856. [Google Scholar] [CrossRef] [PubMed]
  166. Donno, D.; Cavanna, M.; Beccaro, G.L.; Mellano, M.G.; Torello Marinoni, D.; Cerutti, A.K.; Bounous, G. Currants and strawberries as bioactive compound sources: Determination of antioxidant profiles with HPLC-DAD/MS. J. Appl. Bot. Food Qual. 2013, 86, 1–10. [Google Scholar]
  167. Huang, D.; Ou, B.; Hampsch-Woodill, M.; Flanagan, J.A.; Deemer, E.K. Development and validation of oxygen radical absorbance capacity assay for lipophilic antioxidants using randomly methylated β-cyclodextrin as the solubility enhancer. J. Agric. Food Chem. 2002, 50, 1815–1821. [Google Scholar] [CrossRef] [PubMed]
  168. 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] [PubMed]
  169. Benzie, I.F.F.; Strain, J.J. 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. 1998, 299, 15–27. [Google Scholar]
  170. Xie, J.; Schaich, K.M. Re-evaluation of the 2,2-diphenyl-1-picrylhydrazyl free radical (DPPH) assay for antioxidant activity. J. Agric. Food Chem. 2014, 62, 4251–4260. [Google Scholar] [CrossRef] [PubMed]
  171. Arnao, M.B. Some methodological problems in the determination of antioxidant activity using chromogen radicals: A practical case. Trends Food Sci. Technol. 2001, 11, 419–421. [Google Scholar] [CrossRef]
  172. Berker, K.I.; Demirata, B.; Apak, R. Determination of Total Antioxidant Capacity of Lipophilic and Hydrophilic Antioxidants in the Same Solution by Using Ferric-Ferricyanide Assay. Food Anal. Methods 2012, 5, 1150–1158. [Google Scholar] [CrossRef]
  173. 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. [Google Scholar] [CrossRef] [PubMed]
  174. 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. [Google Scholar] [CrossRef] [PubMed]
  175. Mahal, H.S.; Sharma, H.S.; Mukherjee, T. Antioxidant properties of melatonin: A pulse radiolysis study. Free Radic. Biol. Med. 1999, 26, 557–565. [Google Scholar] [CrossRef]
  176. Wardman, P. Reduction Potentials of One Electron Couples Involving Free Radicals in Aqueous Solution. J. Phys. Chem. Ref. Data 1989, 18, 1637–1755. [Google Scholar] [CrossRef]
  177. Paradies, G.; Petrosillo, G.; Paradies, V.; Reiter, R.J.; Ruggiero, F.M. Melatonin, cardiolipin and mitochondrial bioenergetics in health and disease. J. Pineal Res. 2010, 48, 297–310. [Google Scholar] [CrossRef] [PubMed]
  178. Tan, D.X.; Manchester, L.C.; Reiter, R.J.; Qi, W.B.; Karbownik, M.; Calvo, J.R. Significance of melatonin in antioxidative defense system: Reactions and products. Biol. Signals Recept. 2000, 9, 137–159. [Google Scholar] [CrossRef] [PubMed]
  179. Tan, D.X.; Reiter, R.J.; Manchester, L.C.; Yan, M.T.; El-Sawi, M.; Sainz, R.M.; Mayo, J.C.; Kohen, R.; Allegra, M.; Hardeland, R. Chemical and physical properties and potential mechanisms: Melatonin as a broad spectrum antioxidant and free radical scavenger. Curr. Top. Med. Chem. 2002, 2, 181–197. [Google Scholar] [CrossRef] [PubMed]
  180. Hardeland, R.; Reiter, R.J.; Poeggeler, B.; Tan, D.X. The significance of the metabolism of the neurohormone melatonin: Antioxidative protection and formation of bioactive substances. Neurosci. Biobehav. Rev. 1993, 17, 347–357. [Google Scholar] [CrossRef]
  181. Tan, D.X.; Hardeland, R.; Manchester, L.C.; Paredes, S.D.; Korkmaz, A.; Sainz, R.M.; Mayo, J.C.; Fuentes-Broto, L.; Reiter, R.J. The changing biological roles of melatonin during evolution: From an antioxidant to signals of darkness, sexual selection and fitness. Biol. Rev. 2010, 85, 607–623. [Google Scholar] [CrossRef] [PubMed]
  182. Yoshida, M.; Fukuda, A.; Hara, M.; Terada, A.; Kitanaka, Y.; Owada, S. Melatonin prevents the increase in hydroxyl radical-spin trap adduct formation caused by the addition of cisplatin in vitro. Life Sci. 2003, 72, 1773–1780. [Google Scholar] [CrossRef]
  183. Scaiano, J.C. Exploratory laser flash photolysis study of free radical reactions and magnetic field effects in melatonin chemistry. J. Pineal Res. 1995, 19, 189–195. [Google Scholar] [CrossRef] [PubMed]
  184. Reiter, R.J.; Acuña-Castroviejo, D.; Tan, D.X.; Burkhardt, S. Free radical-mediated molecular damage: Mechanisms for the protective actions of melatonin in the central nervous system. Ann. N. Y. Acad. Sci. 2001, 939, 200–215. [Google Scholar] [CrossRef] [PubMed]
  185. Cagnoli, C.M.; Atabay, C.; Kharlamova, E.; Manev, H. Melatonin protects neurons from singlet oxygen-induced apoptosis. J. Pineal Res. 1995, 18, 222–226. [Google Scholar] [CrossRef] [PubMed]
  186. Matuszak, Z.; Bilska, M.A.; Reszkat, K.J.; Chignell, C.F.; Bilski, P. Interaction of Singlet Molecular Oxygen with Melatonin and Related Indoles. Photochem. Photobiol. 2003, 78, 449–455. [Google Scholar] [CrossRef]
  187. Escames, G.; Guerrero, J.M.; Reiter, R.J.; Garcia, J.J.; Munoz-Hoyos, A.; Ortiz, G.G.; Oh, C.S. Melatonin and vitamin E limit nitric oxide-induced lipid peroxidation in rat brain homogenates. Neurosci. Lett. 1997, 230, 147–150. [Google Scholar] [CrossRef]
  188. Siu, A.W.; Ortiz, G.G.; Benitez-King, G.; To, C.H.; Reiter, R.J. Effect of melatonin on the nitric oxide treated retina. Br. J. Ophthalmol. 2004, 88, 1078–1081. [Google Scholar] [CrossRef] [PubMed]
  189. Melchiorri, D.; Reiter, R.J.; Attia, A.M.; Hara, M.; Burgos, A.; Nistico, G. Potent protective effect of melatonin on in vivo paraquat-induced oxidative damage in rats. Life Sci. 1994, 56, 83–89. [Google Scholar] [CrossRef]
  190. Zavodnik, I.B.; Domanski, A.V.; Lapshina, E.A.; Bryszewska, M.; Reiter, R.J. Melatonin directly scavenges free radicals generated in red blood cells and a cell-free system: Chemiluminescence measurements and theoretical calculations. Life Sci. 2006, 79, 391–400. [Google Scholar] [CrossRef] [PubMed]
  191. Gitto, E.; Karbownik, M.; Reiter, R.J.; Xian Tan, D.; Cuzzocrea, S.; Chiurazzi, P.; Cordaro, S.; Corona, G.; Trimarchi, G.; Barberi, I. Effects of melatonin treatment in septic newborns. Pediatr. Res. 2001, 50, 756–760. [Google Scholar] [CrossRef] [PubMed]
  192. Longoni, B.; Salgo, M.G.; Pryor, W.A.; Marchiafava, P.L. Effects of melatonin on lipid peroxidation induced by oxygen radicals. Life Sci. 1998, 62, 853–859. [Google Scholar] [CrossRef]
  193. Taysi, S.; Koc, M.; Büyükokuroǧlu, M.E.; Altinkaynak, K.; Şahin, Y.N. Melatonin reduces lipid peroxidation and nitric oxide during irradiation-induced oxidative injury in the rat liver. J. Pineal Res. 2003, 34, 173–177. [Google Scholar] [CrossRef] [PubMed]
  194. Wakatsuki, A.; Okatani, Y.; Ikenoue, N.; Izumiya, C.; Kaneda, C. Melatonin inhibits oxidative modification of low-density lipoprotein particles in normolipidemic post-menopausal women. J. Pineal Res. 2000, 28, 136–142. [Google Scholar] [CrossRef] [PubMed]
  195. Rosen, J.; Than, N.N.; Koch, D.; Poeggeler, B.; Laatsch, H.; Hardeland, R. Interactions of melatonin and its metabolites with the ABTS cation radical: Extension of the radical scavenger cascade and formation of a novel class of oxidation products, C2-substituted 3-indolinones. J. Pineal Res. 2006, 41, 374–381. [Google Scholar] [CrossRef] [PubMed]
  196. Tan, D.X.; Manchester, L.C.; Terron, M.P.; Flores, L.J.; Reiter, R.J. One molecule, many derivatives: A never-ending interaction of melatonin with reactive oxygen and nitrogen species? J. Pineal Res. 2007, 42, 28–42. [Google Scholar] [CrossRef] [PubMed]
  197. Joshi, N.; Biswas, J.; Nath, C.; Singh, S. Promising Role of Melatonin as Neuroprotectant in Neurodegenerative Pathology. Mol. Neurobiol. 2015, 52, 330–340. [Google Scholar] [CrossRef] [PubMed]
  198. Miller, E.; Morel, A.; Saso, L.; Saluk, J. Melatonin redox activity. Its potential clinical applications in neurodegenerative disorders. Curr. Top. Med. Chem. 2015, 15, 163–169. [Google Scholar] [CrossRef] [PubMed]
  199. Olcese, J.M.; Cao, C.; Mori, T.; Mamcarz, M.B.; Maxwell, A.; Runfeldt, M.J.; Wang, L.; Zhang, C.; Lin, X.; Zhang, G.; et al. Protection against cognitive deficits and markers of neurodegeneration by long-term oral administration of melatonin in a transgenic model of Alzheimer disease. J. Pineal Res. 2009, 47, 82–96. [Google Scholar] [CrossRef] [PubMed]
  200. Dong, W.; Huang, F.; Fan, W.; Cheng, S.; Chen, Y.; Zhang, W.; Shi, H.; He, H. Differential effects of melatonin on amyloid-β peptide 25–35-induced mitochondrial dysfunction in hippocampal neurons at different stages of culture. J. Pineal Res. 2010, 48, 117–125. [Google Scholar] [CrossRef] [PubMed]
  201. Stefanova, N.A.; Maksimova, K.Y.; Kiseleva, E.; Rudnitskaya, E.A.; Muraleva, N.A.; Kolosova, N.G. Melatonin attenuates impairments of structural hippocampal neuroplasticity in OXYS rats during active progression of Alzheimer’s disease-like pathology. J. Pineal Res. 2015, 59, 163–177. [Google Scholar] [CrossRef] [PubMed]
  202. Rudnitskaya, E.A.; Muraleva, N.A.; Maksimova, K.Y.; Kiseleva, E.; Kolosova, N.G.; Stefanova, N.A. Melatonin Attenuates Memory Impairment, Amyloid-β Accumulation, and Neurodegeneration in a Rat Model of Sporadic Alzheimer’s Disease. J. Alzheimer’s Dis. 2015, 47, 103–116. [Google Scholar] [CrossRef] [PubMed]
  203. Ali, T.; Kim, M.O. Melatonin ameliorates amyloid beta-induced memory deficits, tau hyperphosphorylation and neurodegeneration via PI3/Akt/GSk3β pathway in the mouse hippocampus. J. Pineal Res. 2015, 59, 47–59. [Google Scholar] [CrossRef] [PubMed]
  204. Brusco, L.I.; Márquez, M.; Cardinali, D.P. Monozygotic twins with Alzheimer’s disease treated with melatonin: Case report. J. Pineal Res. 1998, 25, 260–263. [Google Scholar] [CrossRef] [PubMed]
  205. Wu, Y.H.; Swaab, D.F. The human pineal gland and melatonin in aging and Alzheimer’s disease. J. Pineal Res. 2005, 38, 145–152. [Google Scholar] [CrossRef] [PubMed]
  206. Carriere, C.H.; Kang, N.H.; Niles, L.P. Chronic low-dose melatonin treatment maintains nigrostriatal integrity in an intrastriatal rotenone model of Parkinson’s disease. Brain Res. 2016, 1633, 115–125. [Google Scholar] [CrossRef] [PubMed]
  207. Yildirim, F.B.; Ozsoy, O.; Tanriover, G.; Kaya, Y.; Ogut, E.; Gemici, B.; Dilmac, S.; Ozkan, A.; Agar, A.; Aslan, M. Mechanism of the beneficial effect of melatonin in experimental Parkinson’s disease. Neurochem. Int. 2014, 79, 1–11. [Google Scholar] [CrossRef] [PubMed]
  208. Bush, S.H.; Lacaze-Masmonteil, N.; McNamara-Kilian, M.T.; MacDonald, A.R.; Tierney, S.; Momoli, F.; Agar, M.; Currow, D.C.; Lawlor, P.G. The preventative role of exogenous melatonin administration to patients with advanced cancer who are at risk of delirium: Study protocol for a randomized controlled trial. Trials 2016, 17, 399. [Google Scholar] [CrossRef] [PubMed]
  209. Innominato, P.F.; Lim, A.S.; Palesh, O.; Clemons, M.; Trudeau, M.; Eisen, A.; Wang, C.; Kiss, A.; Pritchard, K.I.; Bjarnason, G.A. The effect of melatonin on sleep and quality of life in patients with advanced breast cancer. Support. Care Cancer 2016, 24, 1097–1105. [Google Scholar] [CrossRef] [PubMed]
  210. Nooshinfar, E.; Bashash, D.; Safaroghli-Azar, A.; Bayati, S.; Rezaei-Tavirani, M.; Ghaffari, S.H.; Akbari, M.E. Melatonin promotes ATO-induced apoptosis in MCF-7 cells: Proposing novel therapeutic potential for breast cancer. Biomed. Pharmacother. 2016, 83, 456–465. [Google Scholar] [CrossRef] [PubMed]
  211. Sabzichi, M.; Samadi, N.; Mohammadian, J.; Hamishehkar, H.; Akbarzadeh, M.; Molavi, O. Sustained release of melatonin: A novel approach in elevating efficacy of tamoxifen in breast cancer treatment. Colloids Surf. B Biointerfaces 2016, 145, 64–71. [Google Scholar] [CrossRef] [PubMed]
  212. Koşar, P.A.; Nazıroğlu, M.; Övey, İ.S.; Çiğ, B. Synergic Effects of Doxorubicin and Melatonin on Apoptosis and Mitochondrial Oxidative Stress in MCF-7 Breast Cancer Cells: Involvement of TRPV1 Channels. J. Membr. Biol. 2016, 249, 129–140. [Google Scholar] [CrossRef] [PubMed]
  213. Ma, Z.; Yang, Y.; Fan, C.; Han, J.; Wang, D.; Di, S.; Hu, W.; Liu, D.; Li, X.; Reiter, R.J.; et al. Melatonin as a potential anticarcinogen for non-small-cell lung cancer. Oncotarget 2016, 7, 46768–46784. [Google Scholar] [CrossRef] [PubMed]
  214. Akbarzadeh, M.; Nouri, M.; Banekohal, M.V.; Cheraghi, O.; Tajalli, H.; Movassaghpour, A.; Soltani, S.; Cheraghi, H.; Feizy, N.; Montazersaheb, S.; et al. Effects of combination of melatonin and laser irradiation on ovarian cancer cells and endothelial lineage viability. Lasers Med. Sci. 2016, 31, 1565–1572. [Google Scholar] [CrossRef] [PubMed]
  215. Rao, M.V.; Narechania, M.B. The genotoxic effects of anti-cancer drug gossypol on human lymphocytes and its mitigation by melatonin. Drug Chem. Toxicol. 2016, 39, 357–361. [Google Scholar] [CrossRef] [PubMed]
  216. Borin, T.F.; Arbab, A.S.; Gelaleti, G.B.; Ferreira, L.C.; Moschetta, M.G.; Jardim-Perassi, B.V.; Iskander, A.; Varma, N.R.S.; Shankar, A.; Coimbra, V.B.; et al. Melatonin decreases breast cancer metastasis by modulating Rho-associated kinase protein-1 expression. J. Pineal Res. 2016, 60, 3–15. [Google Scholar] [CrossRef] [PubMed]
  217. Paterniti, I.; Cordaro, M.; Esposito, E.; Cuzzocrea, S. The antioxidative property of melatonin against brain ischemia. Expert Rev. Neurother. 2016, 16, 841–848. [Google Scholar] [CrossRef] [PubMed]
  218. Aydemir, S.; Dogan, D.; Kocak, A.; Dilsiz, N. The effect of melatonin on spinal cord after ischemia in rats. Spinal Cord 2016, 54, 360–363. [Google Scholar] [CrossRef] [PubMed]
  219. Bhattacharya, P.; Pandey, A.K.; Paul, S.; Patnaik, R. Melatonin renders neuroprotection by protein kinase C mediated aquaporin-4 inhibition in animal model of focal cerebral ischemia. Life Sci. 2014, 100, 97–109. [Google Scholar] [CrossRef] [PubMed]
  220. Kilic, U.; Yilmaz, B.; Reiter, R.J.; Yüksel, A.; Kilic, E. Effects of memantine and melatonin on signal transduction pathways vascular leakage and brain injury after focal cerebral ischemia in mice. Neuroscience 2013, 237, 268–276. [Google Scholar] [CrossRef] [PubMed]
  221. Yang, Y.; Sun, Y.; Yi, W.; Li, Y.; Fan, C.; Xin, Z.; Jiang, S.; Di, S.; Qu, Y.; Reiter, R.J.; et al. A review of melatonin as a suitable antioxidant against myocardial ischemia-reperfusion injury and clinical heart diseases. J. Pineal Res. 2014, 57, 357–366. [Google Scholar] [CrossRef] [PubMed]
  222. De Souza, A.V.G.; Golim, M.A.; Deffune, E.; Domingues, M.A.C.; de Carvalho, L.R.; Vianna, I.G.; Castiglia, Y.M.M.; Vianna, P.T.G. Evaluation of renal protection from high doses of melatonin in an experimental model of renal ischemia and reperfusion in hyperglycemic rats. Transpl. Proc. 2014, 46, 1591–1593. [Google Scholar] [CrossRef] [PubMed]
  223. Deng, W.S.; Xu, Q.; Liu, Y.; Jiang, C.H.; Zhou, H.; Gu, L. Effects of melatonin on liver function and lipid peroxidation in a rat model of hepatic ischemia/reperfusion injury. Exp. Ther. Med. 2016, 11, 1955–1960. [Google Scholar] [CrossRef] [PubMed]
  224. Li, Y.; Yang, Y.; Feng, Y.; Yan, J.; Fan, C.; Jiang, S.; Qu, Y. A review of melatonin in hepatic ischemia/reperfusion injury and clinical liver disease. Ann. Med. 2014, 46, 503–511. [Google Scholar] [CrossRef] [PubMed]
  225. Kireev, R.; Bitoun, S.; Cuesta, S.; Tejerina, A.; Ibarrola, C.; Moreno, E.; Vara, E.; Tresguerres, J.A.F. Melatonin treatment protects liver of Zucker rats after ischemia/reperfusion by diminishing oxidative stress and apoptosis. Eur. J. Pharmacol. 2013, 701, 185–193. [Google Scholar] [CrossRef] [PubMed]
  226. Yip, H.K.; Chang, Y.C.; Wallace, C.G.; Chang, L.T.; Tsai, T.H.; Chen, Y.L.; Chang, H.W.; Leu, S.; Zhen, Y.Y.; Tsai, C.Y.; et al. Melatonin treatment improves adipose-derived mesenchymal stem cell therapy for acute lung ischemia-reperfusion injury. J. Pineal Res. 2013, 54, 207–221. [Google Scholar] [CrossRef] [PubMed]
  227. Tan, D.X.; Manchester, L.C.; Liu, X.; Rosales-Corral, S.A.; Acuna-Castroviejo, D.; Reiter, R.J. Mitochondria and chloroplasts as the original sites of melatonin synthesis: A hypothesis related to melatonin’s primary function and evolution in eukaryotes. J. Pineal Res. 2013, 54, 127–138. [Google Scholar] [CrossRef] [PubMed]
  228. Slominski, A.T.; Kleszczyński, K.; Semak, I.; Janjetovic, Z.; Żmijewski, M.A.; Kim, T.K.; Slominski, R.M.; Reiter, R.J.; Fischer, T.W. Local melatoninergic system as the protector of skin integrity. Int. J. Mol. Sci. 2014, 15, 17705–17732. [Google Scholar] [CrossRef] [PubMed]
  229. Poeggeler, B.; Reiter, R.J.; Hardeland, R.; Sewerynek, E.; Melchiorri, D.; Barlow-Walden, L.R. Melatonin, a mediator of electron transfer and repair reactions, acts synergistically with the chain-breaking antioxidants ascorbate, trolox and glutathione. Neuroendocrinol. Lett. 1995, 17, 87–92. [Google Scholar]
  230. Gitto, E.; Tan, D.X.; Reiter, R.J.; Karbownik, M.; Manchester, L.C.; Cuzzocrea, S.; Fulia, F.; Barberi, I. Individual and synergistic antioxidative actions of melatonin: Studies with vitamin E, vitamin C, glutathione and desferrrioxamine (desferoxamine) in rat liver homogenates. J. Pharm. Pharmacol. 2001, 53, 1393–1401. [Google Scholar] [CrossRef] [PubMed]
  231. Sliwinski, T.; Rozej, W.; Morawiec-Bajda, A.; Morawiec, Z.; Reiter, R.; Blasiak, J. Protective action of melatonin against oxidative DNA damage-Chemical inactivation versus base-excision repair. Mutat. Res. 2007, 634, 220–227. [Google Scholar] [CrossRef] [PubMed]
  232. Fischer, T.W.; Slominski, A.; Zmijewski, M.A.; Reiter, R.J.; Paus, R. Melatonin as a major skin protectant: From free radical scavenging to DNA damage repair. Exp. Dermatol. 2008, 17, 713–730. [Google Scholar] [CrossRef] [PubMed]
  233. Davanipour, Z.; Poulsen, H.E.; Weimann, A.; Sobel, E. Endogenous melatonin and oxidatively damaged guanine in DNA. BMC Endocr. Disord. 2009, 9, 22. [Google Scholar] [CrossRef] [PubMed]
  234. De Almeida, E.A.; Martinez, G.R.; Klitzke, C.F.; de Medeiros, M.H.G.; Di Mascio, P. Oxidation of melatonin by singlet molecular oxygen (O2(1Δg)) produces N1-acetyl-N2-formyl-5-methoxykynurenine. J. Pineal Res. 2003, 35, 131–137. [Google Scholar] [CrossRef] [PubMed]
  235. Limson, J.; Nyokong, T.; Daya, S. The interaction of melatonin and its precursors with aluminium, cadmium, copper, iron, lead, and zinc: An adsorptive voltammetric study. J. Pineal Res. 1998, 24, 15–21. [Google Scholar] [CrossRef] [PubMed]
  236. Zatta, P.; Tognon, G.; Carampin, P. Melatonin prevents free radical formation due to the interaction between β-amyloid peptides and metal ions [Al(III), Zn(II), Cu(II), Mn(II), Fe(II)]. J. Pineal Res. 2003, 35, 98–103. [Google Scholar] [CrossRef] [PubMed]
  237. Mayo, J.C.; Tan, D.X.; Sainz, R.M.; Natarajan, M.; Lopez-Burillo, S.; Reiter, R.J. Protection against oxidative protein damage induced by metal-catalyzed reaction or alkylperoxyl radicals: Comparative effects of melatonin and other antioxidants. Biochim. Biophys. Acta 2003, 1620, 139–150. [Google Scholar] [CrossRef]
  238. Parmar, P.; Limson, J.; Nyokong, T.; Daya, S. Melatonin protects against copper-mediated free radical damage. J. Pineal Res. 2002, 32, 237–242. [Google Scholar] [CrossRef] [PubMed]
  239. Romero, A.; Ramos, E.; de Los Ríos, C.; Egea, J.; del Pino, J.; Reiter, R.J. A review of metal-catalyzed molecular damage: Protection by melatonin. J. Pineal Res. 2014, 56, 343–370. [Google Scholar] [CrossRef] [PubMed]
  240. Moudache, M.; Colon, M.; Nerín, C.; Zaidi, F. Phenolic content and antioxidant activity of olive by-products and antioxidant film containing olive leaf extract. Food Chem. 2016, 212, 521–527. [Google Scholar] [CrossRef] [PubMed]
  241. Saxena, M.; Saxena, J.; Pradhan, A. Flavonoids and phenolic acids as antioxidants in plants and human health. Int. J. Pharm. Sci. Rev. Res. 2012, 16, 130–134. [Google Scholar]
  242. Fernandez-Panchon, M.S.; Villano, D.; Troncoso, A.M.; Garcia-Parrilla, M.C. Antioxidant activity of phenolic compounds: From in vitro results to in vivo evidence. Crit. Rev. Food Sci. Nutr. 2008, 48, 649–671. [Google Scholar] [CrossRef] [PubMed]
  243. 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. [Google Scholar] [CrossRef] [PubMed]
  244. Cao, L.; Yu, H.; Shao, S.; Wang, S.; Guo, Y. Evaluating the antioxidant capacity of polyphenols with an off-on fluorescence probe and the mechanism study. Anal. Methods 2014, 6, 7149–7153. [Google Scholar] [CrossRef]
  245. Mendoza-Wilson, A.M.; Castro-Arredondo, S.I.; Balandrán-Quintana, R.R. Computational study of the structure-free radical scavenging relationship of procyanidins. Food Chem. 2014, 161, 155–161. [Google Scholar] [CrossRef] [PubMed]
  246. Li, X.; Gao, Y.; Li, F.; Liang, A.; Xu, Z.; Bai, Y.; Mai, W.; Han, L.; Chen, D. Maclurin protects against hydroxyl radical-induced damages to mesenchymal stem cells: Antioxidant evaluation and mechanistic insight. Chem. Biol. Interact. 2014, 219, 221–228. [Google Scholar] [CrossRef] [PubMed]
  247. Wang, G.; Xue, Y.; An, L.; Zheng, Y.; Dou, Y.; Zhang, L.; Liu, Y. Theoretical study on the structural and antioxidant properties of some recently synthesised 2,4,5-trimethoxy chalcones. Food Chem. 2014, 171, 89–97. [Google Scholar] [CrossRef] [PubMed]
  248. Praveena, R.; Sadasivam, K.; Deepha, V.; Sivakumar, R. Antioxidant potential of orientin: A combined experimental and DFT approach. J. Mol. Struct. 2014, 1061, 114–123. [Google Scholar] [CrossRef]
  249. Mikulski, D.; Eder, K.; Molski, M. Quantum-chemical study on relationship between structure and antioxidant properties of hepatoprotective compounds occurring in Cynara scolymus and Silybum marianum. J. Theor. Comput. Chem. 2014, 13, 1450004. [Google Scholar] [CrossRef]
  250. Galano, A.; Martínez, A. Capsaicin, a tasty free radical scavenger: Mechanism of action and kinetics. J. Phys. Chem. B 2012, 116, 1200–1208. [Google Scholar] [CrossRef] [PubMed]
  251. Martínez, A.; Galano, A.; Vargas, R. Free radical scavenger properties of α-mangostin: Thermodynamics and kinetics of HAT and RAF mechanisms. J. Phys. Chem. B 2011, 115, 12591–12598. [Google Scholar] [CrossRef] [PubMed]
  252. Dimitrić Marković, J.M.; Milenković, D.; Amić, D.; Mojović, M.; Pašti, I.; Marković, Z.S. The preferred radical scavenging mechanisms of fisetin and baicalein towards oxygen-centred radicals in polar protic and polar aprotic solvents. RSC Adv. 2014, 4, 32228–32236. [Google Scholar] [CrossRef]
  253. Mazzone, G.; Toscano, M.; Russo, N. Density functional predictions of antioxidant activity and UV spectral features of nasutin A, isonasutin, ellagic acid, and one of its possible derivatives. J. Agric. Food Chem. 2013, 61, 9650–9657. [Google Scholar] [CrossRef] [PubMed]
  254. Xue, Y.; Zheng, Y.; An, L.; Zhang, L.; Qian, Y.; Yu, D.; Gong, X.; Liu, Y. A theoretical study of the structure-radical scavenging activity of hydroxychalcones. Comput. Theor. Chem. 2012, 982, 74–83. [Google Scholar] [CrossRef]
  255. Jeremić, S.; Filipović, N.; Peulić, A.; Marković, Z. Thermodynamical aspect of radical scavenging activity of alizarin and alizarin red S. Theoretical comparative study. Comput. Theor. Chem. 2014, 1047, 15–21. [Google Scholar] [CrossRef]
  256. Xue, Y.; Zheng, Y.; An, L.; Dou, Y.; Liu, Y. Density functional theory study of the structure-antioxidant activity of polyphenolic deoxybenzoins. Food Chem. 2014, 151, 198–206. [Google Scholar] [CrossRef] [PubMed]
  257. Medina, M.E.; Galano, A.; Alvarez-Idaboy, J.R. Theoretical study on the peroxyl radicals scavenging activity of esculetin and its regeneration in aqueous solution. Phys. Chem. Chem. Phys. 2014, 16, 1197–1207. [Google Scholar] [CrossRef] [PubMed]
  258. Marković, Z.; Crossed D Signorović, J.; Dimitrić Marković, J.M.; Živić, M.; Amić, D. Investigation of the radical scavenging potency of hydroxybenzoic acids and their carboxylate anions. Monatsh. Chem. 2014, 145, 953–962. [Google Scholar] [CrossRef]
  259. Pérez-González, A.; Galano, A.; Alvarez-Idaboy, J.R. Dihydroxybenzoic acids as free radical scavengers: Mechanisms, kinetics, and trends in activity. New J. Chem. 2014, 38, 2639–2652. [Google Scholar] [CrossRef]
  260. Medina, M.E.; Iuga, C.; Álvarez-Idaboy, J.R. Antioxidant activity of fraxetin and its regeneration in aqueous media. A density functional theory study. RSC Adv. 2014, 4, 52920–52932. [Google Scholar] [CrossRef]
  261. Caicedo, C.; Iuga, C.; Castañeda-Arriaga, R.; Alvarez-Idaboy, J.R. Antioxidant activity of selected natural polyphenolic compounds from soybean via peroxyl radical scavenging. RSC Adv. 2014, 4, 38918–38930. [Google Scholar] [CrossRef]
  262. 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. [Google Scholar] [CrossRef] [PubMed]
  263. Cordova-Gomez, M.; Galano, A.; Alvarez-Idaboy, J.R. Piceatannol, a better peroxyl radical scavenger than resveratrol. RSC Adv. 2013, 3, 20209–20218. [Google Scholar] [CrossRef]
  264. Shang, Y.J.; Qian, Y.P.; Liu, X.D.; Dai, F.; Shang, X.L.; Jia, W.Q.; Liu, Q.; Fang, J.G.; Zhou, B. Radical-scavenging activity and mechanism of resveratrol-oriented analogues: Influence of the solvent, radical, and substitution. J. Org. Chem. 2009, 74, 5025–5031. [Google Scholar] [CrossRef] [PubMed]
  265. Xue, Y.; Zhang, L.; Li, Y.; Yu, D.; Zheng, Y.; An, L.; Gong, X.; Liu, Y. A DFT study on the structure and radical scavenging activity of newly synthesized hydroxychalcones. J. Phys. Org. Chem. 2013, 26, 240–248. [Google Scholar] [CrossRef]
  266. Marković, Z.; Milenković, D.; Orović, J.; Dimitrić Marković, J.M.; Stepanić, V.; Lučić, B.; Amić, D. Free radical scavenging activity of morin 2′-O-phenoxide anion. Food Chem. 2012, 135, 2070–2077. [Google Scholar] [CrossRef] [PubMed]
  267. Di Meo, F.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Duroux, J.L.; Olivier, Y.; Trouillas, P. Free radical scavenging by natural polyphenols: Atom versus electron transfer. J. Phys. Chem. A 2013, 117, 2082–2092. [Google Scholar] [CrossRef] [PubMed]
  268. Mazzone, G.; Galano, A.; Alvarez-Idaboy, J.R.; Russo, N. Coumarin-Chalcone Hybrids as Peroxyl Radical Scavengers: Kinetics and Mechanisms. J. Chem. Inf. Model. 2016, 56, 662–670. [Google Scholar] [CrossRef] [PubMed]
  269. 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. [Google Scholar] [CrossRef] [PubMed]
  270. Vargas-Sánchez, R.D.; Mendoza-Wilson, A.M.; Torrescano-Urrutia, G.R.; Sánchez-Escalante, A. Antiradical potential of phenolic compounds fingerprints of propolis extracts: DFT approach. Comp. Theor. Chem. 2015, 1066, 7–13. [Google Scholar] [CrossRef]
  271. Kicel, A.; Michel, P.; Owczarek, A.; Marchelak, A.; Zyzelewicz, D.; Budryn, G.; Oracz, J.; Anna Olszewska, M. Phenolic profile and antioxidant potential of leaves from selected Cotoneaster Medik. Species. Molecules 2016, 21, 688. [Google Scholar] [CrossRef] [PubMed]
  272. Medina, M.E.; Iuga, C.; Alvarez-Idaboy, J.R. Antioxidant activity of propyl gallate in aqueous and lipid media: A theoretical study. Phys. Chem. Chem. Phys. 2013, 15, 13137–13146. [Google Scholar] [CrossRef] [PubMed]
  273. Dimitrić Marković, J.M.; Milenković, D.; Amić, D.; Popović-Bijelić, A.; Mojović, M.; Pašti, I.A.; Marković, Z.S. Energy requirements of the reactions of kaempferol and selected radical species in different media: Towards the prediction of the possible radical scavenging mechanisms. Struct. Chem. 2014, 25, 1795–1804. [Google Scholar] [CrossRef]
  274. Inagaki, T.; Yamamoto, T. Critical role of deep hydrogen tunneling to accelerate the antioxidant reaction of ubiquinol and vitamin e. J. Phys. Chem. B 2014, 118, 937–950. [Google Scholar] [CrossRef] [PubMed]
  275. Li, M.; Liu, W.; Peng, C.; Ren, Q.; Lu, W.; Deng, W. A DFT study on reaction of eupatilin with hydroxyl radical in solution. Int. J. Quantum Chem 2013, 113, 966–974. [Google Scholar] [CrossRef]
  276. Amorati, R.; Baschieri, A.; Morroni, G.; Gambino, R.; Valgimigli, L. Peroxyl Radical Reactions in Water Solution: A Gym for Proton-Coupled Electron-Transfer Theories. Chem. Eur. J. 2016, 22, 7924–7934. [Google Scholar] [CrossRef] [PubMed]
  277. Nakayama, T.; Uno, B. Importance of proton-coupled electron transfer from natural phenolic compounds in superoxide scavenging. Chem. Pharm. Bull. 2015, 63, 967–973. [Google Scholar] [CrossRef] [PubMed]
  278. Urbaniak, A.; Szelag, M.; Molski, M. Theoretical investigation of stereochemistry and solvent influence on antioxidant activity of ferulic acid. Comput. Theor. Chem. 2013, 1012, 33–40. [Google Scholar] [CrossRef]
  279. Fifen, J.J.; Nsangou, M.; Dhaouadi, Z.; Motapon, O.; Jaidane, N. Solvent effects on the antioxidant activity of 3,4-dihydroxyphenylpyruvic acid: DFT and TD-DFT studies. Comput. Theor. Chem. 2011, 966, 232–243. [Google Scholar] [CrossRef]
  280. Benayahoum, A.; Amira-Guebailia, H.; Houache, O. Homolytic and heterolytic O-H bond cleavage in trans-resveratrol and some phenantrene analogs: A theoretical study. Comput. Theor. Chem. 2014, 1037, 1–9. [Google Scholar] [CrossRef]
  281. Qian, Y.P.; Shang, Y.J.; Teng, Q.F.; Chang, J.; Fan, G.J.; Wei, X.; Li, R.R.; Li, H.P.; Yao, X.J.; Dai, F.; et al. Hydroxychalcones as potent antioxidants: Structure-activity relationship analysis and mechanism considerations. Food Chem. 2011, 126, 241–248. [Google Scholar] [CrossRef]
  282. Musialik, M.; Kuzmicz, R.; Pawlowski, T.S.; Litwinienko, G. Acidity of hydroxyl groups: An overlooked influence on antiradical properties of flavonoids. J. Org. Chem. 2009, 74, 2699–2709. [Google Scholar] [CrossRef] [PubMed]
  283. Martínez, A.; Hernández-Marin, E.; Galano, A. Xanthones as antioxidants: A theoretical study on the thermodynamics and kinetics of the single electron transfer mechanism. Food Funct. 2012, 3, 442–450. [Google Scholar] [CrossRef] [PubMed]
  284. Najafi, M. On the antioxidant activity of ortho- and meta-substituted indolin-2-one derivatives. Monatsh. Chem. 2014, 145, 291–299. [Google Scholar] [CrossRef]
  285. Najafi, M. On the antioxidant activity of the Ortho and Meta substituted Daidzein derivatives in the gas phase and solvent environment. J. Mex. Chem. Soc. 2014, 58, 36–45. [Google Scholar]
  286. Dorović, J.; Marković, J.M.D.; Stepanić, V.; Begović, N.; Amić, D.; Marković, Z. Influence of different free radicals on scavenging potency of gallic acid. J. Mol. Model. 2014, 20, 2345. [Google Scholar] [CrossRef] [PubMed]
  287. Marković, Z.; Đorović, J.; Dekić, M.; Radulović, M.; Marković, S.; Ilić, M. DFT study of free radical scavenging activity of erodiol. Chem. Pap. 2013, 67, 1453–1461. [Google Scholar] [CrossRef]
  288. Van Wenum, E.; Jurczakowski, R.; Litwinienko, G. Media effects on the mechanism of antioxidant action of silybin and 2,3-dehydrosilybin: Role of the enol group. J. Org. Chem. 2013, 78, 9102–9112. [Google Scholar] [CrossRef] [PubMed]
  289. Farmanzadeh, D.; Najafi, M. Antioxidant activity of aminothiazol hydroxycoumarin derivatives. J. Theor. Comput. Chem. 2013, 12, 1350058. [Google Scholar] [CrossRef]
  290. Lengyel, J.; Rimarčík, J.; Vagánek, A.; Klein, E. On the radical scavenging activity of isoflavones: Thermodynamics of O-H bond cleavage. Phys. Chem. Chem. Phys. 2013, 15, 10895–10903. [Google Scholar] [CrossRef] [PubMed]
  291. Senthil kumar, K.; Kumaresan, R. A DFT study on the structural, electronic properties and radical scavenging mechanisms of calycosin, glycitein, pratensein and prunetin. Comput. Theor. Chem. 2012, 985, 14–22. [Google Scholar] [CrossRef]
  292. 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. [Google Scholar] [CrossRef] [PubMed]
  293. Marković, Z.S.; Dimitrić Marković, J.M.; Milenković, D.; Filipović, N. Mechanistic study of the structure-activity relationship for the free radical scavenging activity of baicalein. J. Mol. Model. 2011, 17, 2575–2584. [Google Scholar] [CrossRef] [PubMed]
  294. Jeremić, S.R.; Šehović, S.F.; Manojlović, N.T.; Marković, Z.S. Antioxidant and free radical scavenging activity of purpurin. Monatsh. Chem. 2012, 143, 427–435. [Google Scholar] [CrossRef]
  295. Mendoza-Sarmiento, G.; Rojas-Hernández, A.; Galano, A.; Gutiérrez, A. A combined experimental-theoretical study of the acid-base behavior of mangiferin: Implications for its antioxidant activity. RSC Adv. 2016, 6, 51171–51182. [Google Scholar] [CrossRef]
  296. Papadopoulos, A.G.; Nenadis, N.; Sigalas, M.P. DFT study of radical scavenging activity of sesame oil lignans and selected in vivo metabolites of sesamin. Comput. Theor. Chem. 2016, 1077, 125–132. [Google Scholar] [CrossRef]
  297. Marković, Z.; Đorović, J.; Petrović, Z.D.; Petrović, V.P.; Simijonović, D. Investigation of the antioxidant and radical scavenging activities of some phenolic Schiff bases with different free radicals. J. Mol. Model. 2015, 21, 293. [Google Scholar] [CrossRef] [PubMed]
  298. Galano, A.; Alvarez-Idaboy, J.R.; Francisco-Márquez, M. Physicochemical insights on the free radical scavenging activity of sesamol: Importance of the acid/base equilibrium. J. Phys. Chem. B 2011, 115, 13101–13109. [Google Scholar] [CrossRef] [PubMed]
  299. Ouchi, A.; Nagaoka, S.I.; Abe, K.; Mukai, K. Kinetic study of the aroxyl radical-scavenging reaction of α-tocopherol in methanol solution: Notable effect of the alkali and alkaline earth metal salts on the reaction rates. J. Phys. Chem. B 2009, 113, 13322–13331. [Google Scholar] [CrossRef] [PubMed]
  300. Marković, Z.S.; Marković, S.; Dimitrić Marković, J.M.; Milenković, D. Structure and reactivity of baicalein radical cation. Int. J. Quantum Chem. 2012, 112, 2009–2017. [Google Scholar] [CrossRef]
  301. Marković, Z.; Amić, D.; Milenković, D.; Dimitrić-Marković, J.M.; Marković, S. Examination of the chemical behavior of the quercetin radical cation towards some bases. Phys. Chem. Chem. Phys. 2013, 15, 7370–7378. [Google Scholar] [CrossRef] [PubMed]
  302. Joshi, R.; Gangabhagirathi, R.; Venu, S.; Adhikari, S.; Mukherjee, T. Antioxidant activity and free radical scavenging reactions of gentisic acid: In Vitro and pulse radiolysis studies. Free Radic. Res. 2012, 46, 11–20. [Google Scholar] [CrossRef] [PubMed]
  303. Nakanishi, I.; Ohkubo, K.; Miyazaki, K.; Hakamata, W.; Urano, S.; Ozawa, T.; Okuda, H.; Fukuzumi, S.; Ikota, N.; Fukuhara, K. A planar catechin analogue having a more negative oxidation potential than (+)-catechin as an electron transfer antioxidant against a peroxyl radical. Chem. Res. Toxicol. 2004, 17, 26–31. [Google Scholar] [CrossRef] [PubMed]
  304. Nakanishi, I.; Shimada, T.; Ohkubo, K.; Manda, S.; Shimizu, T.; Urano, S.; Okuda, H.; Miyata, N.; Ozawa, T.; Anzai, K.; et al. Involvement of electron transfer in the radical-scavenging reaction of resveratrol. Chem. Lett. 2007, 36, 1276–1277. [Google Scholar] [CrossRef]
  305. Sueishi, Y.; Hori, M.; Kita, M.; Kotake, Y. Nitric oxide (NO) scavenging capacity of natural antioxidants. Food Chem. 2011, 129, 866–870. [Google Scholar] [CrossRef] [PubMed]
  306. Diaz-Uribe, C.E.; Vallejo, W.; Castellar, W.; Trilleras, J.; Ortiz, S.; Rodriguez-Serrano, A.; Zarate, X.; Quiroga, J. Novel (E)-1-(pyrrole-2-yl)-3-(aryl)-2-(propen-1-one) derivatives as efficient singlet oxygen quenchers: Kinetics and quantum chemical calculations. RSC Adv. 2015, 5, 71565–71572. [Google Scholar] [CrossRef]
  307. Asano, M.; Iwahashi, H. Caffeic acid inhibits the formation of 7-carboxyheptyl radicals from oleic acid under flavin mononucleotide photosensitization by scavenging singlet oxygen and quenching the excited state of flavin mononucleotide. Molecules 2014, 19, 12486–12499. [Google Scholar] [CrossRef] [PubMed]
  308. Ohara, K.; Doi, K.; Niizaki, Y.; Nagaoka, S.I. A time-resolved luminescence study on singlet oxygen quenching by hydroxycinnamic acids under acidic, neutral and basic conditions. J. Photochem. Photobiol. A Chem. 2012, 249, 1–8. [Google Scholar] [CrossRef]
  309. Mukai, K.; Ishikawa, E.; Ouchi, A.; Nagaoka, S.I.; Suzuki, T.; Izumisawa, K.; Koike, T. Kinetic study of the quenching reaction of singlet oxygen by α-, β-, γ-, δ-tocotrienols, and palm oil and soybean extracts in solution. Biosci. Biotechnol. Biochem. 2014, 78, 2089–2101. [Google Scholar] [CrossRef] [PubMed]
  310. Choi, D.S.; Jung, M.Y. Protective activities of catechins on singlet oxygen induced photooxidation of α-terpinene in methanol: Structure and singlet oxygen quenching activity relationship. Food Sci. Biotechnol. 2013, 22, 249–256. [Google Scholar] [CrossRef]
  311. Jung, M.Y.; Choi, D.S. Electron spin resonance and luminescence spectroscopic observation and kinetic study of chemical and physical singlet oxygen quenching by resveratrol in methanol. J. Agric. Food Chem. 2010, 58, 11888–11895. [Google Scholar] [CrossRef] [PubMed]
  312. Chadha, R.; Mahal, H.S.; Mukherjee, T.; Kapoor, S. Evidence for a possible role of 3-hydroxyanthranilic acid as an antioxidant. J. Phys. Org. Chem. 2009, 22, 349–354. [Google Scholar] [CrossRef]
  313. Sitarek, P.; Skała, E.; Wysokińska, H.; Wielanek, M.; Szemraj, J.; Toma, M.; Śliwiński, T. The effect of leonurus sibiricus plant extracts on stimulating repair and protective activity against oxidative DNA damage in CHO cells and content of phenolic compounds. Oxid. Med. Cell. Longev. 2016, 2016, 5738193. [Google Scholar] [CrossRef] [PubMed]
  314. Chaudhari, G.M.; Mahajan, R.T. Comparative antioxidant activity of twenty traditional Indian medicinal plants and its correlation with total flavonoid and phenolic content. Int. J. Pharm. Sci. Rev. Res. 2015, 30, 105–111. [Google Scholar]
  315. Vasantha Rupasinghe, H.P.; Nair, S.V.G.; Robinson, R.A. Chemopreventive properties of fruit phenolic compounds and their possible mode of actions. Stud. Nat. Prod. Chem. 2014, 42, 229–266. [Google Scholar]
  316. Kabanda, M.M. A theoretical study of the antioxidant properties of phenolic acid amides investigated through the radical-scavenging and metal chelation mechanisms. Eur. Food Res. Technol. 2015, 241, 553–572. [Google Scholar] [CrossRef]
  317. Liu, H.; Cao, J.; Jiang, W. Evaluation and comparison of vitamin C, phenolic compounds, antioxidant properties and metal chelating activity of pulp and peel from selected peach cultivars. Food Sci. Technol. 2015, 63, 1042–1048. [Google Scholar] [CrossRef]
  318. Działo, M.; Mierziak, J.; Korzun, U.; Preisner, M.; Szopa, J.; Kulma, A. The potential of plant phenolics in prevention and therapy of skin disorders. Int. J. Mol. Sci. 2016, 17, 160. [Google Scholar] [CrossRef] [PubMed]
  319. Rangel-Huerta, O.D.; Pastor-Villaescusa, B.; Aguilera, C.M.; Gil, A. A systematic review of the efficacy of bioactive compounds in cardiovascular disease: Phenolic compounds. Nutrients 2015, 7, 5177–5216. [Google Scholar] [CrossRef] [PubMed]
  320. Bulotta, S.; Celano, M.; Lepore, S.M.; Montalcini, T.; Pujia, A.; Russo, D. Beneficial effects of the olive oil phenolic components oleuropein and hydroxytyrosol: Focus on protection against cardiovascular and metabolic diseases. J. Transl. Med. 2014, 12, 219. [Google Scholar] [CrossRef] [PubMed]
  321. Rodriguez-Mateos, A.; Heiss, C.; Borges, G.; Crozier, A. Berry (poly)phenols and cardiovascular health. J. Agric. Food Chem. 2014, 62, 3842–3851. [Google Scholar] [CrossRef] [PubMed]
  322. Panda, V.; Laddha, A.; Nandave, M.; Srinath, S. Dietary Phenolic Acids of Macrotyloma uniflorum (Horse Gram) Protect the Rat Heart Against Isoproterenol-Induced Myocardial Infarction. Phytother. Res. 2016, 1146–1155. [Google Scholar] [CrossRef] [PubMed]
  323. Rodríguez-Morató, J.; Xicota, L.; Fitó, M.; Farré, M.; Dierssen, M.; de La Torre, R. Potential role of olive oil phenolic compounds in the prevention of neurodegenerative diseases. Molecules 2015, 20, 4655–4680. [Google Scholar] [CrossRef] [PubMed]
  324. Rigacci, S.; Stefani, M. Nutraceuticals and amyloid neurodegenerative diseases: A focus on natural phenols. Expert Rev. Neurother. 2014, 15, 41–52. [Google Scholar] [CrossRef] [PubMed]
  325. Ai, G.; Huang, Z.M.; Liu, Q.C.; Han, Y.Q.; Chen, X. The protective effect of total phenolics from Oenanthe Javanica on acute liver failure induced by d-galactosamine. J. Ethnopharmacol. 2016, 186, 53–60. [Google Scholar] [CrossRef] [PubMed]
  326. Razavi-Azarkhiavi, K.; Iranshahy, M.; Sahebkar, A.; Shirani, K.; Karimi, G. The Protective Role of Phenolic Compounds Against Doxorubicin-induced Cardiotoxicity: A Comprehensive Review. Nutr. Cancer 2016, 68, 892–917. [Google Scholar] [CrossRef] [PubMed]
  327. Parkinson, L.; Keast, R. Oleocanthal, a Phenolic Derived from Virgin Olive Oil: A Review of the Beneficial Effects on Inflammatory Disease. Int. J. Mol. Sci. 2014, 15, 12323–12334. [Google Scholar] [CrossRef] [PubMed]
  328. Perron, N.R.; Brumaghim, J.L. A review of the antioxidant mechanisms of polyphenol compounds related to iron binding. Cell Biochem. Biophys. 2009, 53, 75–100. [Google Scholar] [CrossRef] [PubMed]
  329. Scalbert, A.; Williamson, G. Dietary intake and bioavailability of polyphenols. J. Nutr. 2000, 130 (Suppl. S8), 2073S–2085S. [Google Scholar]
  330. D’Archivio, M.; Filesi, C.; Varì, R.; Scazzocchio, B.; Masella, R. Bioavailability of the polyphenols: Status and controversies. Int. J. Mol. Sci. 2010, 11, 1321–1342. [Google Scholar] [CrossRef] [PubMed]
  331. Tan, D.X.; Manchester, L.C.; Esteban-Zubero, E.; Zhou, Z.; Reiter, R.J. Melatonin as a potent and inducible endogenous antioxidant: Synthesis and metabolism. Molecules 2015, 20, 18886–18906. [Google Scholar] [CrossRef] [PubMed]
  332. Fischer, T.W.; Sweatman, T.W.; Semak, I.; Sayre, R.M.; Wortsman, J.; Slominski, A. Constitutive and UV-induced metabolism of melatonin in keratinocytes and cell-free systems. FASEB J. 2006, 20, E897–E907. [Google Scholar] [CrossRef] [PubMed]
  333. Byeon, Y.; Tan, D.X.; Reiter, R.J.; Back, K. Predominance of 2-hydroxymelatonin over melatonin in plants. J. Pineal Res. 2015, 59, 448–454. [Google Scholar] [CrossRef] [PubMed]
  334. Ximenes, V.F.; Padovan, C.Z.; Carvalho, D.A.; Fernandes, J.R. Oxidation of melatonin by taurine chloramine. J. Pineal Res. 2010, 49, 115–122. [Google Scholar] [CrossRef] [PubMed]
  335. Watts, S.W.; Morrison, S.F.; Davis, R.P.; Barman, S.M. Serotonin and blood pressure regulation. Pharmacol. Rev. 2012, 64, 359–388. [Google Scholar] [CrossRef] [PubMed]
  336. Turner, E.H.; Blackwell, A.D. 5-Hydroxytryptophan plus SSRIs for interferon-induced depression: Synergistic mechanisms for normalizing synaptic serotonin. Med. Hypotheses 2005, 65, 138–144. [Google Scholar] [CrossRef] [PubMed]
  337. Birdsall, T.C. 5-Hydroxytryptophan: A Clinically-Effective Serotonin Precursor. Altern. Med. Rev. 1998, 3, 271–280. [Google Scholar] [PubMed]
  338. Turner, E.H.; Loftis, J.M.; Blackwell, A.D. Serotonin a la carte: Supplementation with the serotonin precursor 5-hydroxytryptophan. Pharmacol. Ther. 2006, 109, 325–338. [Google Scholar] [CrossRef] [PubMed]
  339. Kema, I.P.; De Vries, E.G.E.; Muskiet, F.A.J. Clinical chemistry of serotonin and metabolites. J. Chromatogr. B Biomed. Sci. Appl. 2000, 747, 33–48. [Google Scholar] [CrossRef]
  340. González-Gómez, D.; Lozano, M.; Fernández-León, M.F.; Ayuso, M.C.; Bernalte, M.J.; Rodríguez, A.B. Detection and quantification of melatonin and serotonin in eight Sweet Cherry cultivars (Prunus avium L.). Eur. Food Res. Technol. 2009, 229, 223–229. [Google Scholar] [CrossRef]
  341. Ramakrishna, A.; Giridhar, P.; Sankar, K.U.; Ravishankar, G.A. Melatonin and serotonin profiles in beans of Coffea species. J. Pineal Res. 2012, 52, 470–476. [Google Scholar] [CrossRef] [PubMed]
  342. Ly, D.; Kang, K.; Choi, J.Y.; Ishihara, A.; Back, K.; Lee, S.G. HPLC analysis of serotonin, tryptamine, tyramine, and the hydroxycinnamic acid amides of serotonin and tyramine in food vegetables. J. Med. Food 2008, 11, 385–389. [Google Scholar] [CrossRef] [PubMed]
  343. Kang, S.; Kang, K.; Lee, K.; Back, K. Characterization of tryptamine 5-hydroxylase and serotonin synthesis in rice plants. Plant Cell Rep. 2007, 26, 2009–2015. [Google Scholar] [CrossRef] [PubMed]
  344. Olivier, B. Serotonin: A never-ending story. Eur. J. Pharmacol. 2015, 753, 2–18. [Google Scholar] [CrossRef] [PubMed]
  345. Gershon, M.D. 5-Hydroxytryptamine (serotonin) in the gastrointestinal tract. Curr. Opin. Endocrinol. Diabetes Obes. 2013, 20, 14–21. [Google Scholar] [CrossRef] [PubMed]
  346. Nausheen, F.; Mohsin, H.; Lakhan, S.E. Neurotransmitters in hiccups. SpringerPlus 2016, 5, 1357. [Google Scholar] [CrossRef] [PubMed]
  347. Byeon, Y.; Back, K. Melatonin synthesis in rice seedlings in vivo is enhanced at high temperatures and under dark conditions due to increased serotonin N-acetyltransferase and N-acetylserotonin methyltransferase activities. J. Pineal Res. 2014, 56, 189–195. [Google Scholar] [CrossRef] [PubMed]
  348. Shi, H.; Chan, Z. The cysteine2/histidine2-type transcription factor ZINC FINGER of ARABIDOPSIS THALIANA 6-activated C-REPEAT-BINDING FACTOR pathway is essential for melatonin-mediated freezing stress resistance in Arabidopsis. J. Pineal Res. 2014, 57, 185–191. [Google Scholar] [CrossRef] [PubMed]
  349. Byeon, Y.; Lee, H.Y.; Back, K. Chloroplastic and cytoplasmic overexpression of sheep serotonin N-acetyltransferase in transgenic rice plants is associated with low melatonin production despite high enzyme activity. J. Pineal Res. 2015, 58, 461–469. [Google Scholar] [CrossRef] [PubMed]
  350. Lee, H.Y.; Byeon, Y.; Tan, D.X.; Reiter, R.J.; Back, K. Arabidopsis serotonin N-acetyltransferase knockout mutant plants exhibit decreased melatonin and salicylic acid levels resulting in susceptibility to an avirulent pathogen. J. Pineal Res. 2015, 58, 291–299. [Google Scholar] [CrossRef] [PubMed]
  351. Young, I.M.; Leone, R.M.; Francis, P.; Stovell, P.; Silman, R.E. Melatonin is metabolized to N-acetyl serotonin and 6-hydroxymelatonin in man. J. Clin. Endocrinol. Metab. 1985, 60, 114–119. [Google Scholar] [CrossRef] [PubMed]
  352. Dellegar, S.M.; Murphy, S.A.; Bourne, A.E.; Dicesare, J.C.; Purser, G.H. Identification of the factors affecting the rate of deactivation of hypochlorous acid by melatonin. Biochem. Biophys. Res. Commun. 1999, 257, 431–439. [Google Scholar] [CrossRef] [PubMed]
  353. Agozzino, P.; Avellone, G.; Bongiorno, D.; Ceraulo, L.; Filizzola, F.; Natoli, M.C.; Livrea, M.A.; Tesoriere, L. Melatonin: Structural characterization of its non-enzymatic mono-oxygenate metabolite. J. Pineal Res. 2003, 35, 269–275. [Google Scholar] [CrossRef] [PubMed]
  354. Horstman, J.A.; Wrona, M.Z.; Dryhurst, G. Further insights into the reaction of melatonin with hydroxyl radical. Bioorg. Chem. 2002, 30, 371–382. [Google Scholar] [CrossRef]
  355. Semak, I.; Naumova, M.; Korik, E.; Terekhovich, V.; Wortsman, J.; Slominski, A. A novel metabolic pathway of melatonin: Oxidation by cytochrome C. Biochemistry 2005, 44, 9300–9307. [Google Scholar] [CrossRef] [PubMed]
  356. Byeon, Y.; Back, K. Molecular cloning of melatonin 2-hydroxylase responsible for 2-hydroxymelatonin production in rice (Oryza sativa). J. Pineal Res. 2015, 58, 343–351. [Google Scholar] [CrossRef] [PubMed]
  357. Lahiri, D.K.; Ge, Y.W.; Sharman, E.H.; Bondy, S.C. Age-related changes in serum melatonin in mice: Higher levels of combined melatonin and 6-hydroxymelatonin sulfate in the cerebral cortex than serum, heart, liver and kidney tissues. J. Pineal Res. 2004, 36, 217–223. [Google Scholar] [CrossRef] [PubMed]
  358. Kim, T.K.; Lin, Z.; Tidwell, W.J.; Li, W.; Slominski, A.T. Melatonin and its metabolites accumulate in the human epidermis in vivo and inhibit proliferation and tyrosinase activity in epidermal melanocytes in vitro. Mol. Cell. Endocrinol. 2015, 404, 1–8. [Google Scholar] [CrossRef] [PubMed]
  359. Kim, T.K.; Kleszczynśki, K.; Janjetovic, Z.; Sweatman, T.; Lin, Z.; Li, W.; Reiter, R.J.; Fischer, T.W.; Slominski, A.T. Metabolism of melatonin and biological activity of intermediates of melatoninergic pathway in human skin cells. FASEB J. 2013, 27, 2742–2755. [Google Scholar] [CrossRef] [PubMed]
  360. Wiechmann, A.F.; Sherry, D.M. Role of Melatonin and its Receptors in the Vertebrate Retina. Int. Rev. Cell Mol. Biol. 2013, 300, 211–242. [Google Scholar] [PubMed]
  361. Pinato, L.; da Silveira Cruz-Machado, S.; Franco, D.G.; Campos, L.M.G.; Cecon, E.; Fernandes, P.A.C.M.; Bittencourt, J.C.; Markus, R.P. Selective protection of the cerebellum against intracerebroventricular LPS is mediated by local melatonin synthesis. Brain Struct. Funct. 2013, 220, 1–14. [Google Scholar] [CrossRef] [PubMed]
  362. Cruz, M.H.C.; Leal, C.L.V.; Cruz, J.F.; Tan, D.X.; Reiter, R.J. Essential actions of melatonin in protecting the ovary from oxidative damage. Theriogenology 2014, 82, 925–932. [Google Scholar] [CrossRef] [PubMed]
  363. Acuña-Castroviejo, D.; Escames, G.; Venegas, C.; Díaz-Casado, M.E.; Lima-Cabello, E.; López, L.C.; Rosales-Corral, S.; Tan, D.X.; Reiter, R.J. Extrapineal melatonin: Sources, regulation, and potential functions. Cell. Mol. Life Sci. 2014, 71, 2997–3025. [Google Scholar] [CrossRef] [PubMed]
  364. Reiter, R.J.; Tan, D.X.; Zhou, Z.; Cruz, M.H.C.; Fuentes-Broto, L.; Galano, A. Phytomelatonin: Assisting plants to survive and thrive. Molecules 2015, 20, 7396–7437. [Google Scholar] [CrossRef] [PubMed]
  365. Pan, L.; McKain, B.W.; Madan-Khetarpal, S.; McGuire, M.; Diler, R.S.; Perel, J.M.; Vockley, J.; Brent, D.A. GTP-cyclohydrolase deficiency responsive to sapropterin and 5-HTP supplementation: Relief of treatment-refractory depression and suicidal behaviour. BMJ Case Rep. 2011, 2011, bcr0320113927. [Google Scholar] [PubMed]
  366. Dell’Osso, L.; Carmassi, C.; Mucci, F.; Marazziti, D. Depression, serotonin and tryptophan. Curr. Pharm. Des. 2016, 22, 949–954. [Google Scholar] [CrossRef] [PubMed]
  367. Abdala-Valencia, H.; Berdnikovs, S.; McCary, C.A.; Urick, D.; Mahadevia, R.; Marchese, M.E.; Swartz, K.; Wright, L.; Mutlu, G.M.; Cook-Mills, J.M. Inhibition of allergic inflammation by supplementation with 5-hydroxytryptophan. Am. J. Physiol. Lung Cell Mol. Physiol. 2012, 303, L642–L660. [Google Scholar] [CrossRef] [PubMed]
  368. Zhang, H.; Zhao, H.; Yang, X.; Xue, Q.; Cotten, J.F.; Feng, H.J. 5-Hydroxytryptophan, a precursor for serotonin synthesis, reduces seizure-induced respiratory arrest. Epilepsia 2016, 57, 1228–1235. [Google Scholar] [CrossRef] [PubMed]
  369. Bocchio, M.; McHugh, S.B.; Bannerman, D.M.; Sharp, T.; Capogna, M. Serotonin, amygdala and fear: Assembling the puzzle. Front. Neural Circuits 2016, 10, 24. [Google Scholar] [CrossRef] [PubMed]
  370. Lesch, K.P.; Waider, J. Serotonin in the Modulation of Neural Plasticity and Networks: Implications for Neurodevelopmental Disorders. Neuron 2012, 76, 175–191. [Google Scholar] [CrossRef] [PubMed]
  371. Patrick, R.P.; Ames, B.N. Vitamin D and the omega-3 fatty acids control serotonin synthesis and action, part 2: Relevance for ADHD, bipolar disorder, schizophrenia, and impulsive behavior. FASEB J. 2015, 29, 2207–2222. [Google Scholar] [CrossRef] [PubMed]
  372. Lesch, K.P.; Araragi, N.; Waider, J.; van den Hove, D.; Gutknecht, L. Targeting brain serotonin synthesis: Insights into neurodevelopmental disorders with long-term outcomes related to negative emotionality, aggression and antisocial behaviour. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2012, 367, 2426–2443. [Google Scholar] [CrossRef] [PubMed]
  373. Sanfey, A.G. Social decision-making: Insights from game theory and neuroscience. Science 2007, 318, 598–602. [Google Scholar] [CrossRef] [PubMed]
  374. Way, B.M.; Laćan, G.; Fairbanks, L.A.; Melega, W.P. Architectonic distribution of the serotonin transporter within the orbitofrontal cortex of the vervet monkey. Neuroscience 2007, 148, 937–948. [Google Scholar] [CrossRef] [PubMed]
  375. Pagan, C.; Delorme, R.; Callebert, J.; Goubran-Botros, H.; Amsellem, F.; Drouot, X.; Boudebesse, C.; Dudal, K.L.; Ngo-Nguyen, N.; Laouamri, H.; et al. The serotonin-N-acetylserotonin-melatonin pathway as a biomarker for autism spectrum disorders. Transl. Psychiatry 2014, 4, e479. [Google Scholar] [CrossRef] [PubMed]
  376. Lee, H.Y.; Byeon, Y.; Back, K. Melatonin as a signal molecule triggering defense responses against pathogen attack in Arabidopsis and tobacco. J. Pineal Res. 2014, 57, 262–268. [Google Scholar] [CrossRef] [PubMed]
  377. Zhang, N.; Sun, Q.; Zhang, H.; Cao, Y.; Weeda, S.; Ren, S.; Guo, Y.D. Roles of melatonin in abiotic stress resistance in plants. J. Exp. Bot. 2015, 66, 647–656. [Google Scholar] [CrossRef] [PubMed]
  378. Janjetovic, Z.; Nahmias, Z.P.; Hanna, S.; Jarrett, S.G.; Kim, T.K.; Reiter, R.J.; Slominski, A.T. Melatonin and its metabolites ameliorate ultraviolet B-induced damage in human epidermal keratinocytes. J. Pineal Res. 2014, 57, 90–102. [Google Scholar] [CrossRef] [PubMed]
  379. Calvo, J.R.; Reiter, R.J.; Garca, J.J.; Ortiz, G.G.; Tan, D.X.; Karbownik, M. Characterization of the protective effects of melatonin and related indoles against α-naphthylisothiocyanate-induced liver injury in rats. J. Cell. Biochem. 2001, 80, 461–470. [Google Scholar] [CrossRef]
  380. García, J.J.; Reiter, R.J.; Karbownik, M.; Calvo, J.R.; Ortiz, G.G.; Tan, D.X.; Martínez-Ballarín, E.; Acua-Castroviejo, D. N-acetylserotonin suppresses hepatic microsomal membrane rigidity associated with lipid peroxidation. Eur. J. Pharmacol. 2001, 428, 169–175. [Google Scholar] [CrossRef]
  381. García, J.J.; Lõpez-Pingarrõn, L.; Almeida-Souza, P.; Tres, A.; Escudero, P.; García-Gil, F.A.; Tan, D.X.; Reiter, R.J.; Ramírez, J.M.; Bernal-Pérez, M. Protective effects of melatonin in reducing oxidative stress and in preserving the fluidity of biological membranes: A review. J. Pineal Res. 2014, 56, 225–237. [Google Scholar] [CrossRef] [PubMed]
  382. García, J.J.; Reiter, R.J.; Guerrero, J.M.; Escames, G.; Yu, B.P.; Oh, C.S.; Muñoz-Hoyos, A. Melatonin prevents changes in microsomal membrane fluidity during induced lipid peroxidation. FEBS Lett. 1997, 408, 297–300. [Google Scholar] [CrossRef]
  383. Anderson, G.; Rodriguez, M. Multiple sclerosis: The role of melatonin and N-acetylserotonin. Mult. Scler. Relat. Disord. 2015, 4, 112–123. [Google Scholar] [CrossRef] [PubMed]
  384. Yu, S.; Zheng, J.; Jiang, Z.; Shi, C.; Li, J.; Du, X.; Wang, H.; Jiang, J.; Wang, X. Protective effect of N-acetylserotonin against acute hepatic ischemia-reperfusion injury in mice. Int. J. Mol. Sci. 2013, 14, 17680–17693. [Google Scholar] [CrossRef] [PubMed]
  385. Galano, A.; Medina, M.E.; Tan, D.X.; Reiter, R.J. Melatonin and its metabolites as copper chelating agents and their role in inhibiting oxidative stress: A physicochemical analysis. J. Pineal Res. 2015, 58, 107–116. [Google Scholar] [CrossRef] [PubMed]
  386. Christen, S.; Peterhans, E.; Stocker, R. Antioxidant activities of some tryptophan metabolites: Possible implication for inflammatory diseases. Proc. Natl. Acad. Sci. USA 1990, 87, 2506–2510. [Google Scholar] [CrossRef] [PubMed]
  387. Cadenas, E.; Simic, M.G.; Sies, H. Antioxidant activity of 5-hydroxytryptophan, 5-hydroxyindole, and dopa against microsomal lipid peroxidation and its dependence on vitamin E. Free Radic. Res. 1989, 6, 11–17. [Google Scholar] [CrossRef]
  388. Millán-Plano, S.; Piedrafita, E.; Miana-Mena, F.J.; Fuentes-Broto, L.; Martínez-Ballarín, E.; López-Pingarrón, L.; Sáenz, M.A.; García, J.J. Melatonin and structurally-related compounds protect synaptosomal membranes from free radical damage. Int. J. Mol. Sci. 2010, 11, 312–328. [Google Scholar] [CrossRef] [PubMed]
  389. Weber, O.A.; Simeon, V. Tryptamine, 5-hydroxytryptamine and 5-hydroxytryptophan complexes of proton and some divalent metal ions. J. Inorg. Nucl. Chem. 1971, 33, 2097–2101. [Google Scholar] [CrossRef]
  390. Lysek, N.; Kinscherf, R.; Claus, R.; Lindel, T. l-5-hydroxytryptophan: Antioxidant and anti-apoptotic principle of the intertidal sponge Hymeniacidon heliophila. Z. Naturforsch. C 2003, 58, 568–572. [Google Scholar] [CrossRef] [PubMed]
  391. Bae, S.J.; Lee, J.S.; Kim, J.M.; Lee, E.K.; Han, Y.K.; Kim, H.J.; Choi, J.; Ha, Y.M.; No, J.K.; Kim, Y.H.; et al. 5-hydroxytrytophan inhibits tert-butylhydroperoxide (t-BHP)-induced oxidative damage via the suppression of reactive species (RS) and nuclear factor-κB (NF-κB) activation on human fibroblast. J. Agric. Food Chem. 2010, 58, 6387–6394. [Google Scholar] [CrossRef] [PubMed]
  392. Reyes-Gonzales, M.C.; Fuentes-Broto, L.; Martínez-Ballarín, E.; Miana-Mena, F.J.; Berzosa, C.; García-Gil, F.A.; Aranda, M.; García, J.J. Effects of tryptophan and 5-hydroxytryptophan on the hepatic cell membrane rigidity due to oxidative stress. J. Membr. Biol. 2009, 231, 93–99. [Google Scholar] [CrossRef] [PubMed]
  393. Yang, T.H.; Hsu, P.Y.; Meng, M.; Su, C.C. Supplement of 5-hydroxytryptophan before induction suppresses inflammation and collagen-induced arthritis. Arthritis Res. Ther. 2015, 17, 364. [Google Scholar] [CrossRef] [PubMed]
  394. Keithahn, C.; Lerchl, A. 5-Hydroxytryptophan is a more potent in vitro hydroxyl radical scavenger than melatonin or vitamin C. J. Pineal Res. 2005, 38, 62–66. [Google Scholar] [CrossRef] [PubMed]
  395. Herraiz, T.; Galisteo, J. Endogenous and dietary indoles: A class of antioxidants and radical scavengers in the ABTS assay. Free Radic. Res. 2004, 38, 323–331. [Google Scholar] [CrossRef] [PubMed]
  396. Siddhuraju, P.; Becker, K. Studies on antioxidant activities of mucuna seed (Mucuna pruriens var utilis) extract and various non-protein amino/imino acids through in vitro models. J. Sci. Food Agric. 2003, 83, 1517–1524. [Google Scholar] [CrossRef]
  397. Noda, Y.; Mori, A.; Liburdy, R.; Packer, L. Melatonin and its precursors scavenge nitric oxide. J. Pineal Res. 1999, 27, 159–163. [Google Scholar] [CrossRef] [PubMed]
  398. Derlacz, R.A.; Sliwinska, M.; Piekutowska, A.; Winiarska, K.; Drozak, J.; Bryla, J. Melatonin is more effective than taurine and 5-hydroxytryptophan against hyperglycemia-induced kidney-cortex tubules injury. J. Pineal Res. 2007, 42, 203–209. [Google Scholar] [CrossRef] [PubMed]
  399. Daniels, W.M.U.; van Rensburg, S.J.; van Zyl, J.M.; van Der Walt, B.J.; Taljaard, J.J.F. Free radical scavenging effects of melatonin and serotonin: Possible mechanism. Neuroreport 1996, 7, 1593–1596. [Google Scholar] [CrossRef] [PubMed]
  400. Huether, G.; Fettkötter, I.; Keilhoff, G.; Wolf, G. Serotonin acts as a radical scavenger and is oxidized to a dimer during the respiratory burst of activated microglia. J. Neurochem. 1997, 69, 2096–2101. [Google Scholar] [CrossRef] [PubMed]
  401. Schuff-Werner, P.; Splettstoesser, W. Antioxidative properties of serotonin and the bactericidal function of polymorphonuclear phagocytes. Adv. Exp. Med. Biol. 2000, 467, 321–325. [Google Scholar]
  402. Sarikaya, S.B.Ö.; Gülçin, I. Radical scavenging and antioxidant capacity of serotonin. Curr. Bioact. Comp. 2013, 9, 143–152. [Google Scholar] [CrossRef]
  403. Gülçin, I. Measurement of antioxidant ability of melatonin and serotonin by the DMPD and CUPRAC methods as trolox equivalent. J. Enzyme Inhib. Med. Chem. 2008, 23, 871–876. [Google Scholar] [CrossRef] [PubMed]
  404. Betten, Å.; Dahlgren, C.; Hermodsson, S.; Hellstrand, K. Serotonin protects NK cells against oxidatively induced functional inhibition and apoptosis. J. Leukoc. Biol. 2001, 70, 65–72. [Google Scholar] [PubMed]
  405. Sang Soo, H.; Dong Hyun, K.; Suk Ha, L.; Yun Sang, K.; Chung Soo, L. Antioxidant effects of serotonin and l-DOPA on oxidative damages of brain synaptosomes. Korean J. Physiol. Pharmacol. 1999, 3, 147–155. [Google Scholar]
  406. Longoni, B.; Pryor, W.A.; Marchiafava, P. Inhibition of lipid peroxidation by N-acetylserotonin and its role in retinal physiology. Biochem. Biophys. Res. Commun. 1997, 233, 778–780. [Google Scholar] [CrossRef] [PubMed]
  407. Wölfler, A.; Abuja, P.M.; Schauenstein, K.; Liebmann, P.M. N-acetylserotonin is a better extra- and intracellular antioxidant than melatonin. FEBS Lett. 1999, 449, 206–210. [Google Scholar] [CrossRef]
  408. Zhou, H.; Wang, J.; Jiang, J.; Stavrovskaya, I.G.; Li, M.; Li, W.; Wu, Q.; Zhang, X.; Luo, C.; Zhou, S.; et al. N-acetyl-serotonin offers neuroprotection through inhibiting mitochondrial death pathways and autophagic activation in experimental models of ischemic injury. J. Neurosci. 2014, 34, 2967–2978. [Google Scholar] [CrossRef] [PubMed]
  409. Jiang, J.; Yu, S.; Jiang, Z.; Liang, C.; Yu, W.; Li, J.; Du, X.; Wang, H.; Gao, X.; Wang, X. N-acetyl-serotonin protects HepG2 cells from oxidative stress injury induced by hydrogen peroxide. Oxid. Med. Cell. Longev. 2014, 2014, 310504. [Google Scholar] [CrossRef] [PubMed]
  410. Aguiar, L.M.; Macedo, D.S.; de Freitas, R.M.; de Albuquerque Oliveira, A.; Vasconcelos, S.M.M.; de Sousa, F.C.F.; de Barros Viana, G.S. Protective effects of N-acetylserotonin against 6-hydroxydopamine-induced neurotoxicity. Life Sci. 2005, 76, 2193–2202. [Google Scholar] [CrossRef] [PubMed]
  411. Bachurin, S.; Oxenkrug, G.F.; Lermontova, N.; Afanasiev, A.; Beznosko, B.; Vankin, G.; Shevtzova, E.; Mukhina, T.; Serkova, T. N-acetylserotonin, melatonin and their derivatives improve cognition and protect against β-amyloid-induced neurotoxicity. Ann. N. Y. Acad. Sci. 1999, 890, 155–166. [Google Scholar] [CrossRef] [PubMed]
  412. Oxenkrug, G.; Requintina, P.; Bachurin, S. Antioxidant and antiaging activity of N-acetylserotonin and melatonin in the in vivo models. Ann. N. Y. Acad. Sci. 2001, 939, 190–199. [Google Scholar] [CrossRef] [PubMed]
  413. Tosini, G.; Ye, K.; Iuvone, P.M. N-acetylserotonin: Neuroprotection, neurogenesis, and the sleepy brain. Neuroscientist 2012, 18, 645–653. [Google Scholar] [CrossRef] [PubMed]
  414. Chan, T.Y.; Tang, P.L. Characterization of the antioxidant effects of melatonin and related indoleamines in vitro. J. Pineal Res. 1996, 20, 187–191. [Google Scholar] [CrossRef] [PubMed]
  415. Lezoualc’h, F.; Sparapani, M.; Behl, C. N-acetyl-serotonin (normelatonin) and melatonin protect neurons against oxidative challenges and suppress the activity of the transcription factor NF-κB. J. Pineal Res. 1998, 24, 168–178. [Google Scholar] [CrossRef] [PubMed]
  416. Tang, G.Y.; Ip, A.K.; Siu, A.W. Pinoline and N-acetylserotonin reduce glutamate-induced lipid peroxidation in retinal homogenates. Neurosci. Lett. 2007, 412, 191–194. [Google Scholar] [CrossRef] [PubMed]
  417. Seeger, H.; Mueck, A.O.; Lippert, T.H. Effect of melatonin and metabolites on copper-mediated oxidation of low density lipoprotein. Br. J. Clin. Pharmacol. 1997, 44, 283–284. [Google Scholar] [CrossRef]
  418. Gozzo, A.; Lesieur, D.; Duriez, P.; Fruchart, J.C.; Teissier, E. Structure-activity relationships in a series of melatonin analogues with the low-density lipoprotein oxidation model. Free Radic. Biol. Med. 1999, 26, 1538–1543. [Google Scholar] [CrossRef]
  419. Siu, A.W.; Cheung, J.P.; To, C.H.; Chan, E.K.; Chan, J.K.; Cheung, J.C. N-acetyl-serotonin reduces copper(I) ion-induced lipid peroxidation in bovine retinal homogenates. Acta Ophthalmol. Scand. 2001, 79, 69–71. [Google Scholar] [CrossRef] [PubMed]
  420. Qi, W.; Reiter, R.J.; Tan, D.X.; Manchester, L.C.; Siu, A.W.; Garcia, J.J. Increased levels of oxidatively damaged DNA induced by chromium(III) and H2O2: Protection by melatonin and related molecules. J. Pineal Res. 2000, 29, 54–61. [Google Scholar] [CrossRef] [PubMed]
  421. Guajardo, M.H.; Terrasa, A.M.; Catalá, A. Protective effect of indoleamines on in vitro ascorbate-Fe 2+dependent lipid peroxidation of rod outer segment membranes of bovine retina. J. Pineal Res. 2003, 35, 276–282. [Google Scholar] [CrossRef] [PubMed]
  422. Karbownik, M.; Gitto, E.; Lewiñski, A.; Reiter, R.J. Relative efficacies of indole antioxidants in reducing autoxidation and iron-induced lipid peroxidation in hamster testes. J. Cell. Biochem. 2001, 81, 693–699. [Google Scholar] [CrossRef] [PubMed]
  423. Oxenkrug, G. Antioxidant effects of N-acetylserotonin: Possible mechanisms and clinical implications. Ann. N. Y. Acad. Sci. 2005, 1053, 334–347. [Google Scholar] [CrossRef] [PubMed]
  424. Pierrefiche, G.; Topall, G.; Courboin, G.; Henriet, I.; Laborit, H. Antioxidant activity of melatonin in mice. Res. Commun. Chem. Pathol. Pharmacol. 1993, 80, 211–224. [Google Scholar] [PubMed]
  425. Maharaj, D.S.; Walker, R.B.; Glass, B.D.; Daya, S. 6-Hydroxymelatonin protects against cyanide induced oxidative stress in rat brain homogenates. J. Chem. Neuroanat. 2003, 26, 103–107. [Google Scholar] [CrossRef]
  426. Maharaj, D.S.; Limson, J.L.; Daya, S. 6-Hydroxymelatonin converts Fe(III) to Fe(II) and reduces iron-induced lipid peroxidation. Life Sci. 2003, 72, 1367–1375. [Google Scholar] [CrossRef]
  427. Maharaj, D.S.; Maharaj, H.; Antunes, E.M.; Maree, D.M.; Nyokong, T.; Glass, B.D.; Daya, S. 6-Hydroxymelatonin protects against quinolinic-acid-induced oxidative neurotoxicity in the rat hippocampus. J. Pharm. Pharmacol. 2005, 57, 877–881. [Google Scholar] [CrossRef] [PubMed]
  428. Maharaj, D.S.; Anoopkumar-Dukie, S.; Glass, B.D.; Antunes, E.M.; Lack, B.; Walker, R.B.; Daya, S. The identification of the UV degradants of melatonin and their ability to scavenge free radicals. J. Pineal Res. 2002, 32, 257–261. [Google Scholar] [CrossRef] [PubMed]
  429. Maharaj, D.S.; Maharaj, H.; Daya, S.; Glass, B.D. Melatonin and 6-hydroxymelatonin protect against iron-induced neurotoxicity. J. Neurochem. 2006, 96, 78–81. [Google Scholar] [CrossRef] [PubMed]
  430. 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. [Google Scholar] [CrossRef] [PubMed]
  431. Zhao, Y.; Schultz, N.E.; Truhlar, D.G. Design of density functionals by combining the method of constraint satisfaction with parametrization for thermochemistry, thermochemical kinetics, and noncovalent interactions. J. Chem. Theory Comput. 2006, 2, 364–382. [Google Scholar] [CrossRef] [PubMed]
  432. Galano, A.; Pérez-González, A. On the free radical scavenging mechanism of protocatechuic acid, regeneration of the catechol group in aqueous solution. Theor. Chem. Acc. 2012, 131, 1–13. [Google Scholar] [CrossRef]
  433. Galano, A.; Francisco Marquez, M.; Pérez-González, A. Ellagic acid: An unusually versatile protector against oxidative stress. Chem. Res. Toxicol. 2014, 27, 904–918. [Google Scholar] [CrossRef] [PubMed]
  434. Álvarez-Diduk, R.; Galano, A. Adrenaline and noradrenaline: Protectors against oxidative stress or molecular targets? J. Phys. Chem. B 2015, 119, 3479–3491. [Google Scholar] [CrossRef] [PubMed]
  435. 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 Del. Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef]
  436. 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. [Google Scholar] [CrossRef] [PubMed]
  437. Sverdlov, R.L.; Khrishchanovich, A.V.; Brinkevich, S.D.; Shadyro, O.I. Interaction of tryptophan and its derivatives with oxygen-and nitrogen-centered radicals. High Energy Chem. 2015, 49, 83–91. [Google Scholar] [CrossRef]
  438. Sverdlov, R.L.; Brinkevich, S.D.; Shadyro, O.I. Interaction of tryptophan and related compounds with oxygen- and carbon-centered radicals. Free Radic. Res. 2014, 48, 1200–1205. [Google Scholar] [CrossRef] [PubMed]
  439. Rodriguez-Naranjo, M.I.; Moyá, M.L.; Cantos-Villar, E.; Garcia-Parrilla, M.C. Comparative evaluation of the antioxidant activity of melatonin and related indoles. J. Food Compos. Anal. 2012, 28, 16–22. [Google Scholar] [CrossRef]
  440. Najafi, M.; Farmanzadeh, D.; Klein, E.; Zahedi, M. A Theoretical study on the enthalpies of homolytic and heterolytic N-H bond cleavage in substituted melatonins in the gas-phase and aqueous solution. Acta Chim. Slov. 2013, 60, 43–55. [Google Scholar] [PubMed]
  441. Pérez-González, A.; Alvarez-Idaboy, J.R.; Galano, A. Free-radical scavenging by tryptophan and its metabolites through electron transfer based processes. J. Mol. Model. 2015, 21, 213. [Google Scholar] [CrossRef] [PubMed]
  442. Álvarez-Diduk, R.; Galano, A.; Tan, D.X.; Reiter, R.J. The key role of the sequential proton loss electron transfer mechanism on the free radical scavenging activity of some melatonin-related compounds. Theor. Chem. Acc. 2016, 135, 1–5. [Google Scholar] [CrossRef]
  443. Frankel, E.N.; Huang, S.-W.; Kanner, J.; German, J.B. Interfacial Phenomena in the Evaluation of Antioxidants: Bulk Oils vs. Emulsions. J. Agric. Food Chem. 1994, 42, 1054–1059. [Google Scholar] [CrossRef]
  444. Frankel, E.N.; Huang, S.-W.; Aeschbach, R.; Prior, E. Antioxidant Activity of a Rosemary Extract and Its Constituents, Carnosic Acid, Carnosol, and Rosmarinic Acid, in Bulk Oil and Oil-in-Water Emulsion. J. Agric. Food Chem. 1996, 44, 131–135. [Google Scholar] [CrossRef]
  445. Simko, F.; Bednarova, K.R.; Krajcirovicova, K.; Hrenak, J.; Celec, P.; Kamodyova, N.; Gajdosechova, L.; Zorad, S.; Adamcova, M. Melatonin reduces cardiac remodeling and improves survival in rats with isoproterenol-induced heart failure. J. Pineal Res. 2014, 57, 177–184. [Google Scholar] [CrossRef] [PubMed]
  446. Fagali, N.; Catalá, A. The antioxidant behaviour of melatonin and structural analogues during lipid peroxidation depends not only on their functional groups but also on the assay system. Biochem. Biophys. Res. Commun. 2012, 423, 873–877. [Google Scholar] [CrossRef] [PubMed]
  447. Halliwell, B.; Murcia, M.A.; Chirico, S.; Aruoma, O.I. Free radicals and antioxidants in food and in vivo: What they do and how they work. Crit. Rev. Food Sci. Nutr. 1995, 35, 7–20. [Google Scholar] [CrossRef] [PubMed]
  448. Halliwell, B.; Whiteman, M. Measuring reactive species and oxidative damage in vivo and in cell culture: How should you do it and what do the results mean? Br. J. Pharmacol. 2004, 142, 231–255. [Google Scholar] [CrossRef] [PubMed]
  449. Kawanishi, S.; Sakurai, H. Differential anti-lipid peroxidative activity of melatonin. Naturwissenschaften 2002, 89, 31–33. [Google Scholar] [CrossRef] [PubMed]
  450. Gaikwad, P.; Naik, G.H.; Priyadarsini, K.I.; Mohan, H.; Rao, B.S.M. Radiation induced oxidation of hydroxy indoles by NO2 and Br•−2-radicals: Effect of pH. J. Phys. Org. Chem. 2011, 24, 657–662. [Google Scholar] [CrossRef]
  451. Matuszak, Z.; Reszka, K.J.; Chignell, C.F. Reaction of melatonin and related indoles with hydroxyl radicals: EPR and spin trapping investigations. Free Radic. Biol. Med. 1997, 23, 367–372. [Google Scholar] [CrossRef]
  452. Roberts, J.E.; Hu, D.N.; Wishart, J.F. Pulse radiolysis studies of melatonin and chloromelatonin. J. Photochem. Photobiol. B Biol. 1998, 42, 125–132. [Google Scholar] [CrossRef]
  453. Stasica, P.; Ulanski, P.; Rosiak, J.M. Reactions of melatonin with radicals in deoxygenated aqueous solution. J. Radioanal. Nucl. Chem. 1998, 232, 107–113. [Google Scholar] [CrossRef]
  454. Poeggeler, B.; Reiter, R.J.; Hardeland, R.; Tan, D.X.; Barlow-Walden, L.R. Melatonin and structurally-related, endogenous indoles act as potent electron donors and radical scavengers in vitro. Redox Rep. 1996, 2, 179–184. [Google Scholar] [CrossRef] [PubMed]
  455. Chyan, Y.J.; Poeggeler, B.; Omar, R.A.; Chain, D.G.; Frangione, B.; Ghiso, J.; Pappolla, M.A. Potent neuroprotective properties against the Alzheimer β-amyloid by an endogenous melatonin-related indole structure, indole-3-propionic acid. J. Biol. Chem. 1999, 274, 21937–21942. [Google Scholar] [CrossRef] [PubMed]
  456. 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. [Google Scholar] [CrossRef] [PubMed]
  457. Dorfman, L.M.; Adams, G.E. Reactivity of the Hydroxyl Radical in Aqueous Solutions; U.S. Department of Commerce, National Bureau of Standards: Washington, DC, USA, 1973.
  458. Wilson, W.E. A critical review of the gas phase reaction kinetics of the hydroxyl radical. J. Phys. Chem. Ref. Data 1972, 1, 535–573. [Google Scholar] [CrossRef]
  459. Gulcin, I.; Buyukokuroglu, M.E.; Kufrevioglu, O.I. Metal chelating and hydrogen peroxide scavenging effects of melatonin. J. Pineal Res. 2003, 34, 278–281. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of melatonin (MLT).
Figure 1. Structure of melatonin (MLT).
Molecules 21 01442 g001
Figure 2. Phenolic melatonin-related compounds analyzed in this review.
Figure 2. Phenolic melatonin-related compounds analyzed in this review.
Molecules 21 01442 g002
Figure 3. Schematic representation of the metabolic route connecting the naturally occurring phenolic melatonin-related compounds. Enzymes include cytochrome P450 (CP450), horseradish peroxidase (HRP), indoleamine 2,3-dioxygenase (IDO), eosinophil peroxidase (EPO), myeloperoxidase (MPO) and melatonin 2-hydroxylase (M2H). Oxidants include ROS and RNS.
Figure 3. Schematic representation of the metabolic route connecting the naturally occurring phenolic melatonin-related compounds. Enzymes include cytochrome P450 (CP450), horseradish peroxidase (HRP), indoleamine 2,3-dioxygenase (IDO), eosinophil peroxidase (EPO), myeloperoxidase (MPO) and melatonin 2-hydroxylase (M2H). Oxidants include ROS and RNS.
Molecules 21 01442 g003
Figure 4. Relative rate constants (k) for the reactions with HOO (a) in lipid solution; (b) in aqueous solution, at pH = 7.4. The x-axis values correspond to log(ki/kTrolox). Data on Trolox, ascorbic acid, resveratrol, caffeine, propyl gallate, and caffeic acid was taken from references [120,173,263,272,292,456].
Figure 4. Relative rate constants (k) for the reactions with HOO (a) in lipid solution; (b) in aqueous solution, at pH = 7.4. The x-axis values correspond to log(ki/kTrolox). Data on Trolox, ascorbic acid, resveratrol, caffeine, propyl gallate, and caffeic acid was taken from references [120,173,263,272,292,456].
Molecules 21 01442 g004
Figure 5. MLT, and related compounds, grouped according to their main AOC. A question mark next to the acronym of a compound means that further investigations are needed to confirm the proposed classification, thus its position can eventually change. AFMK = N1-acetyl-N2-formyl-5-methoxykynuramine, c3-OHM = cyclic 3-hydroxymelatonin.
Figure 5. MLT, and related compounds, grouped according to their main AOC. A question mark next to the acronym of a compound means that further investigations are needed to confirm the proposed classification, thus its position can eventually change. AFMK = N1-acetyl-N2-formyl-5-methoxykynuramine, c3-OHM = cyclic 3-hydroxymelatonin.
Molecules 21 01442 g005
Table 1. Some of the most common oxidants in living organisms.
Table 1. Some of the most common oxidants in living organisms.
ROSRNSRSS
NameAcronym/FormulaeNameAcronym/FormulaeNameAcronym/Formulae
superoxide radical anionO2•−peroxynitriteONOOthiyl radicalsRS
hydroxyl radicalOHnitric oxideNOsulfenic acidsRSOH
alkoxyl radicalsROnitrogen dioxideNO2disulfide-S-oxidesRS(O)2SR
peroxyl radicalsROO
Table 2. Most common reaction mechanisms involved in the primary AOC of chemical antioxidants.
Table 2. Most common reaction mechanisms involved in the primary AOC of chemical antioxidants.
NameAcronymChemical Reaction *
Radical Adduct FormationRAFHnAtx + R → [HnAtx-R]
Single Electron TransferSETHnAtx + R → HnAtx+• + R (I)
HnAtx + R → HnAtx−• + R+ (II)
Hydrogen Atom TransferHATHnAtx + R → Hn−1Atx + HR
Proton Coupled Electron TransferPCETHnAtx + R → Hn−1Atx + HR
Sequential Proton Loss Electron TransferSPLETHnAtx → Hn−1Atx + H+
Hn−1Atx + R → Hn−1Atx + R
Sequential Electron Proton TransferSEPTHnAtx + R → Hn−1Atx•+ + R
Hn−1Atx•+ → Hn−1Atx + H+
* HnAtx = chemical antioxidant.
Table 3. Rate constants (k) of the reviewed compounds as primary antioxidants, i.e., as FR scavengers. MLT has been including for comparison purposes.
Table 3. Rate constants (k) of the reviewed compounds as primary antioxidants, i.e., as FR scavengers. MLT has been including for comparison purposes.
AntioxidantFRk (M−1·s−1)AssayMain SolventReference
5-HTPOH1.2 × 1010Experimentalnot specified[449]
5-HTPNO29.0 × 105Experimentalwater, pH = 5[450]
5-HTPNO25.6 × 107Experimentalwater, pH = 9[450]
5-HTPO2•−1.2 × 104Experimentalnot specified[449]
5-HTPDPPH~7 × 10−1Experimentalmethanol[439]
NASDPPH~2 × 100Experimentalmethanol[439]
NASHOO6.70 × 104Theoreticallipid[43]
NASHOO1.17 × 106Theoreticalwater, pH = 7.4[53]
6-HMLTOH1.1 × 1010Experimentalwater, pH = 7[451]
6-HMLTO2•−2.7 × 104Experimentalnot specified[449]
6-HMLTHOO5.81 × 103Theoreticallipid[43]
6-HMLTHOO3.62 × 106Theoreticalwater, pH = 7.4[53]
4-HMLTHOO5.48 × 105TheoreticallipidThis work
4-HMLTHOO6.21 × 106Theoreticalwater, pH = 7.4This work
C1HOO4.79 × 106Theoreticalwater, pH = 7.4[53]
C2HOO6.79 × 106Theoreticalwater, pH = 7.4[53]
C3HOO2.54 × 106Theoreticalwater, pH = 7.4[53]
C4HOO1.61 × 105Theoreticalwater, pH = 7.4[53]
C5HOO1.89 × 105Theoreticalwater, pH = 7.4[53]
MLTOH2.23 × 1010Theoreticalbenzene[131]
MLTOH2.57 × 1010Experimentalwater(a)
MLTOOH3.11 × 102Theoreticalbenzene[131]
MLTOOH1.99 × 101Theoreticalwater, pH = 7.4[131]
MLTO2•−<1.0 × 104Experimentalwater[452]
MLTNO23.7 × 106Experimentalwater, pH = 7[175]
(a) Average from the values reported in references [131,175,451,452,453,454,455].

Share and Cite

MDPI and ACS Style

Galano, A.; Castañeda-Arriaga, R.; Pérez-González, A.; Tan, D.-X.; Reiter, R.J. Phenolic Melatonin-Related Compounds: Their Role as Chemical Protectors against Oxidative Stress. Molecules 2016, 21, 1442. https://doi.org/10.3390/molecules21111442

AMA Style

Galano A, Castañeda-Arriaga R, Pérez-González A, Tan D-X, Reiter RJ. Phenolic Melatonin-Related Compounds: Their Role as Chemical Protectors against Oxidative Stress. Molecules. 2016; 21(11):1442. https://doi.org/10.3390/molecules21111442

Chicago/Turabian Style

Galano, Annia, Romina Castañeda-Arriaga, Adriana Pérez-González, Dun-Xian Tan, and Russel J. Reiter. 2016. "Phenolic Melatonin-Related Compounds: Their Role as Chemical Protectors against Oxidative Stress" Molecules 21, no. 11: 1442. https://doi.org/10.3390/molecules21111442

Article Metrics

Back to TopTop