Next Article in Journal
Recent Advances in Understanding the Control of Secretory Proteins by the Unfolded Protein Response in Plants
Next Article in Special Issue
Melatonin Effects on Hard Tissues: Bone and Tooth
Previous Article in Journal
Simple and Rapid Synthesis of Magnetite/Hydroxyapatite Composites for Hyperthermia Treatments via a Mechanochemical Route
Previous Article in Special Issue
Role of Melatonin in Schizophrenia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuroprotective Effect of Melatonin: A Novel Therapy against Perinatal Hypoxia-Ischemia

Department of Cell Biology and Histology, School of Medicine and Dentistry, University of the Basque Country, Barrio Sarriena s/n, Leioa 48940, Bizkaia, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2013, 14(5), 9379-9395; https://doi.org/10.3390/ijms14059379
Submission received: 28 February 2013 / Revised: 15 April 2013 / Accepted: 16 April 2013 / Published: 29 April 2013
(This article belongs to the Special Issue Advances in the Research of Melatonin)

Abstract

:
One of the most common causes of mortality and morbidity in children is perinatal hypoxia-ischemia (HI). In spite of the advances in neonatology, its incidence is not diminishing, generating a pediatric population that will require an extended amount of chronic care throughout their lifetime. For this reason, new and more effective neuroprotective strategies are urgently required, in order to minimize as much as possible the neurological consequences of this encephalopathy. In this sense, interest has grown in the neuroprotective possibilities of melatonin, as this hormone may help to maintain cell survival through the modulation of a wide range of physiological functions. Although some of the mechanisms by which melatonin is neuroprotective after neonatal asphyxia remain a subject of investigation, this review tries to summarize some of the most recent advances related with its use as a therapeutic drug against perinatal hypoxic-ischemic brain injury, supporting the high interest in this indoleamine as a future feasible strategy for cerebral asphyctic events.

Graphical Abstract

1. Introduction

The main function of the pineal gland in all species is to transduce information concerning light-dark cycles to body physiology, particularly for organization of body rhythms via its main hormone melatonin [1]. Based on the ability of melatonin (N-acetyl-5-methoxytryptamine) and its metabolites to scavenge a wide variety of free radicals (FR), it is not surprising to consider it as one of its most important functions in living organisms leading to protect them from oxidative stress [2,3]. Acting as a direct scavenger, this neurohormone is able to remove FR, such as singlet oxygen, superoxide anion radical, hydroperoxide, hydroxyl radical and the lipid peroxide radical [2,4]. Moreover, a single melatonin molecule may generate products in a scavenger cascade, which may collectively eliminate up to ten FR [4]. Melatonin can develop indirect antioxidant actions through the improvement of the mitochondrial efficiency [5], the stimulation of the gene expression and the activation of some of the most important antioxidant enzymes, including superoxide dismutase (SOD), catalase, glucose-6-phosphate dehydrogenase, glutathione reductase and glutathione peroxidase [6] and also with the strengthening of the antioxidant effect of substances, like glutathione, vitamin E and vitamin C [7].
The brain is particularly sensitive to FR damage due to its high utilization of oxygen, its relatively poorly developed antioxidant defense and its high amount of easily oxidizable fatty acids. Thus, the use of melatonin as pharmacological agent against neurodegenerative disorders, such as Huntington’s disease, Alzheimer’s disease and Parkinsonism and also against ischemic brain injury/stroke, has been extensively evaluated. With an incidence of 2–6/1,000 term births [8], perinatal hypoxia-ischemia (HI) remains the single most important cause of brain injury in the newborn, leading to death or lifelong disability [9,10]. Despite the improvements in perinatal care, significant neurological sequelae can occur in as many as 50%–75% of these asphyctic children, which may suffer from long-term neurological consequences, such as cerebral palsy, mental retardation and epilepsy [1113]. Asphyxia is also associated with attention deficits and hyperactivity in children and adolescents [14,15]. During the last decade, melatonin has started to be considered as an attractive option in order to minimize as much as possible the neurological sequelae from hypoxic-ischemic brain injury. It easily crosses both the placental and the blood-brain barrier, reaching subcellular compartments with a low toxicity and high efficacy [1618], and even at high supra-physiological concentrations, there appear to be no adverse side-effects [19], making it a relatively-safe therapy that could be administered to babies.

2. Brain Protection

Following cerebral asphyxia, HI starts out a multi-faceted cascade of events that ultimately causes cell death and often damages the whole brain [20]. Everything begins when the reduction in oxygen and blood supply induces a decrease in oxidative phosphorylation and the neonate’s brain converts to anaerobic metabolism in an effort to sustain functional ability. Anaerobic metabolism leads to a rapid depletion of ATP, accumulation of lactic acid and failure of ion pumps, resulting in a massive entry of sodium, calcium and water into the cells. Afterwards, multiple and diverse downstream biochemical reactions aggravate the pathogenesis of hypoxic-ischemic brain damage, being the most important, among others, the production of reactive oxygen species leading to oxidative stress, the massive increase in free cytosolic calcium concentrations and the drop in mitochondrial function triggering the activation of apoptotic pathways, DNA fragmentation and cell death. After ischemic brain injury/stroke, melatonin has showed a remarkable capacity to reduce infarct volume and/or inhibit neuronal cell death after in different mammalian species and using different experimental models [2133].
Melatonin administration after neonatal HI has been shown to reduce infarct volume both administered before or after the injury [3436]. Indeed, melatonin was able to decrease sensorimotor asymmetry and learning deficits, thus protecting the pups from the long-term consequences of neonatal asphyxia [34]. While virtually every cell is affected by asphyxia, they do not respond in the same way during HI, being neurons the most sensitive cells to the lack of oxygen and showing a selective vulnerability [3739]. Histological analysis demonstrated an increase in the number of morphologically well preserved neurons in melatonin-treated animals in the CA1, CA2–CA3 areas and dentate gyrus of the hippocampus and parietal cortex when compared with the hypoxic-ischemic group [33,40,41] (Figure 1). Even though astrocytes are more resistant than neurons to oxygen deprivation, their death may give rise to a new wage of neuronal death due to their role in the maintenance of the homeostatic state of neurons. Therefore, astrocytes can modulate in a significant manner the extension and degree of severity of the damage [42,43], either conferring neuroprotection by scavenging reactive oxygen species and also assisting with reconstruction from brain injury [44] or leading to deficiencies in the myelination processes, neuronal signaling impairment and an increase the inflammatory response [45,46]. Melatonin has demonstrated to reduce the expression of the glial fibrillary acidic protein [33] (Figure 1), whose accumulation is related with the creation of new astrocytic processes and reactive gliosis [43]. In addition to neurons, oligodendrocytes are particularly vulnerable to asphyxia, affecting myelination that gives rise to white matter lesions and damaging gray matter oligodendrocyte progenitors [47]. An abnormal decrease in the expression of myelin basic protein leading to myelination deficit is considered hallmark of inflammation-associated diffuse white matter damage [48,49]. In this sense, several groups have suggested that melatonin may be of therapeutic value in ameliorating hypoxic-ischemic damage to the developing white matter through normalization of the myelination process [33,5052] (Figure 1).

3. Antioxidant

The high incidence of hypoxic-ischemic brain lesions in newborns can be partly attributed to the fact that the developing brain is especially vulnerable to oxidative stress. After neonatal asphyxia, reperfusion brings about the overproduction of FR leading to oxidative stress, as the antioxidant capacity of immature neurons is easily overwhelmed by hypoxia-induced reactive oxygen species. The newborn brain is especially vulnerable to oxidative imbalance, due to its increased fatty acid content, higher concentrations of free iron, high rates of oxygen consumption, low concentrations of antioxidant, an imbalance of antioxidant enzymes, as for example catalase CuZn-SOD-1, mitochondrial SOD-2 and glutathione peroxidase and oxygen-induced vasoconstriction, leading to reduced brain perfusion, among others [5355].
When membrane lipoproteins and polyunsaturated fatty acids suffer attacks from FR, many oxygenated compounds, particularly aldehydes, such as malondialdehyde (MDA), are produced. Thus, the evaluation of lipid peroxidation is a useful tool to evaluate oxidative stress leading to brain damage. Melatonin has been able to abolish lipid peroxidation in late-gestation fetal sheep in respond to umbilical cord occlusion [56] and to avoid the rise in MDA induced by hypoxia in rat pups [57] and asphyxiated human newborns [58]. Deferoxamine-chelatable free iron, isoprostanes, neuroprostanes and neurofurans are also quantitative biomarkers of oxidative damage [5961], and after melatonin administration, their levels were significantly lower than those in hypoxic-ischemic rats [62,63]. These results were similar to those observed in fetal sheep after umbilical cord occlusion, where the production of 8-isoprostanes was attenuated [64]. On the other hand, melatonin may prevent protein oxidation in the brain tissue of hypoxic neonatal rats [65], as terminal products of protein exposure to FR are considered reliable markers of the degree of protein damage in oxidative stress [66]. Additionally, the activity of the antioxidative enzyme catalase is maintained [57], hydroxyl formation reduced [56] and nitrite/nitrate levels reduced [58] in different animals models subjected to hypoxia.

4. Anti-Apoptotic

Under pathophysiological conditions, one of the most important key regulators of apoptotic cell death is mitochondrial impairment, as the disruption of its membrane integrity and loss of membrane potential can determinate cell survival by overproduction of reactive oxygen species, abnormal calcium homeostasis and release of apoptotic proteins. In several studies demonstrating anti-apoptotic actions, melatonin prevented cytochrome c release [32,67,68], reduced or blocked caspase-1 and caspase-3 activation [32,67,6973], increased the expression of anti-apoptotic proteins Bcl-2 [71,74,75] and Bcl-xL[70], diminished Bad [31,72] and Bax [71] pro-apoptotic proteins, inhibited poly-ADP-ribose-polymerase cleavage [72], avoided mitochondrial permeability transition pore opening, thus counteracting the collapse of the mitochondrial membrane potential [67,69], and decreased the number of TUNEL-positive cells/DNA breaks [31,32,72,7578]. In the central nervous system, melatonin can also generate anti-excitatory effects on neurons through the modulation of gamma-aminobutyric acid and glutamate receptors [79,80], inducing a decrease in cytosolic calcium concentrations [81,82].
Shortly after a hypoxic-ischemic event, reactive oxygen species overproduction can start out a harmful multi-faceted cascade that includes lipid peroxidation, protein oxidation and DNA fragmentation, ultimately damaging vital cellular components as nucleic acids, cell membranes and mitochondria, resulting in subsequent cell death in the immature brain [8385]. In this regard, we have recently shown that HI can develop a widespread increase in reactive oxygen species after perinatal asphyxia, an overproduction correlated with the number of, as well as with the distribution of apoptotic cells [86].
Melatonin may be an effective prophylactic agent for use in late pregnancy to protect against mitochondrial-induced cell death after a hypoxic-ischemic event at birth. Given to pregnant rats, it prevented oxidative mitochondria damage after ischemia-reperfusion in premature fetal rat brain [87] by means of the maintenance of the number of intact mitochondria and the respiratory control index (an indicator of mitochondrial respiratory activity), as well as the reduction in thiobarbituric acid-reactive substances concentration (a marker of oxidative stress) [40,41]. Hutton et al. studied caspase-3 activation and fractin in order to evaluate the anti-apoptotic effect of melatonin in a model of birth asphyxia in the spiny mouse, showing lower levels after its administration [88]. Accordingly, Fu et al. demonstrated the inhibition of caspase-3 activation, the induction of Bcl-2 expression and the increase Bcl-2/Bax ratio in a model of hypoxia in vitro [89]. Melatonin administration not only generates a neuroprotective effect when administered before the hypoxic-ischemic event, but also when given after the onset of the injury. Using the terminal deoxynucleotidyl transferase dUTP nick end labeling method (TUNEL) to detect DNA fragmentation and apoptotic figures, melatonin-treated neonatal animals have shown a reduction in TUNEL-positive cells per unit area in neonatal sheep [64] and in rats [33,35,36].
As shown above, melatonin can exert a wide range of antiapoptotic effects, mainly targeting mitochondria, but it can also enhance cell survival pathways leading to cell rescue. For instance, protection from cerebral ischemic injury was attributed to the maintenance of signaling via the MAP kinase pathway, leading to the prevention Bad dephosphorylation [31]. Furthermore, melatonin can target PI3K/Akt pathway [70,72,9092], mTOR [93] or the forkhead transcription factor, pAFX [91], and also restore JNK1/2 and ERK 1/2 phosphorylated levels [72,90], thereby preventing the proapoptotic actions of the dephosphorylated proteins.

5. Anti-Inflammatory

In recent years, the use of melatonin has started to be considered as another meaningful tool against inflammatory response in an effort to improve the clinical course of illnesses, which have an inflammatory etiology. The strategy mediated by melatonin and its main metabolites 6-hydroxy, N1-acetyl-N2-formyl-5-metho, N1-acetyl-5-methoxykynuramine and cyclic 3-hydroxy melatonin, encompasses the downregulation of some inflammation-related molecules, such as cytokines interleukin-6, interleukin-8 and tumor necrosis factor-α [94100], 5-lipoxygenase [101], cyclooxygenase [102,103] and prostaglandin [104106], an important reduction of nitric oxide (NO) and MDA levels [107] and also the inhibition of neuronal (nNOS) [108111] and inducible (iNOS) nitric oxide synthases [111,112]. Even though NO participates in diverse processes acting as a physiological messenger, an excess in its concentrations can induce energy depletion, liberation of excitotoxic amino acids and a high ability to react with other FR. Downregulation of nNOS may contribute to the maintenance of electron transport chain function [109,110,113115], thus protecting from NO-mediated mitochondrial impairment and cell damage [116119]. From its part, counteraction to iNOS can avoid lipid peroxidation, shifts the glutathione redox state and boosts energy efficiency and ATP production in mitochondria [118,120,121]. Regarding the field of ischemic brain injury, melatonin and its metabolites have been able to reverse the inflammatory response and edema after stroke suppressing the production of inflammatory cytokines [102,122,123], reducing NOS [124], preventing the translocation of NF-κB to the nucleus [125,126] and decreasing cyclooxygenase-2 gene expression [127], molecular changes correlated with a reduction in the size of brain infarcts.
After perinatal asphyxia, FR also stimulate ischemic cells to secrete inflammatory cytokines and chemokines, which in turn can generate a wide variety of cytotoxic agents, including more cytokines, matrix metalloproteases, NO and more reactive oxygen species. These molecules can dismantle both blood brain barrier and extracellular matrix, allowing the blood, soluble elements and peripheral inflammatory cells to penetrate the brain, resulting in the exacerbation of the damage. Melatonin may be beneficial, as it reduces NO production, vascular endothelial growth factor concentration and, hence, vascular permeability that results increased after hypoxic exposure [128]. Prophylactic maternal treatment with melatonin has also demonstrated a reduction in central nervous system inflammation, by limiting macrophage infiltration and glial cell activation in a model of birth asphyxia in the spiny mouse [88]. Indeed, a reduced number of ED1 positive cells, a marker of activated microglia-macrophages, was found in neonatal rats treated with melatonin when comparing with pups without treatment [63].

6. Conclusions

Nowadays, there are convincing evidences demonstrating that melatonin treatment is highly effective against hypoxic-ischemic brain injury in different animal models by reducing infarct volume and neuronal loss, minimizing lipid and protein peroxidation, blocking some apoptotic pathways, inhibiting FR production and decreasing inflammation. Melatonin supplementation, which has a benign safety profile, may help to reduce complications in the neonatal period that are associated with short gestation [129] and has demonstrated not only neuroprotective actions against HI in animal models, but also in preliminary clinical trials [57,130132].
Nevertheless, the complexity of neonatal hypoxic-ischemic pathophysiology determines that successful neuroprotection could be achieved only by multi-therapeutic approaches and optimizing therapy for neonatal brain injury will require capitalizing on multiple pathways, which prevent cell death. The use of synergic strategies, such as the association between hypothermia and other therapeutic drugs, may lead to a larger neuroprotective effect on the brain thus improving the neonatal outcome. In this regard, Robertson et al. have recently shown that melatonin administration to newborn piglets augments hypothermic neuroprotection by improving cerebral energy metabolism and by reducing brain damage [133].
Melatonin’s protective actions include not only its direct free radical scavenging, but also the interaction of its receptors and several yet-undefined functions, so the mechanisms underlying its neuroprotective benefits are not yet fully elucidated. Moreover, its variable oral absorption and rapid metabolization [134136], the search for an appropriate dosage to obtain an antioxidant effect without desensitize melatonin receptors and its different pharmacokinetic profile when comparing preterm infants with adults (the half-life of melatonin in neonates is approximately 15 h, while in adults, it is around 45–60 min) [137], highlight the work that needs to be done before melatonin comes into clinical practice in a neonatal or pediatric critical care unit.

Conflict of Interest

The authors declare no conflict of interest.

References

  1. Arendt, J. Melatonin and the Mammalian Pineal Gland; Chapman & Hall: London, UK, 1995. [Google Scholar]
  2. 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]
  3. 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. Camb. Philos. Soc 2010, 85, 607–623. [Google Scholar]
  4. 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]
  5. Acuna-Castroviejo, D.; Martin, M.; Macias, M.; Escames, G.; Leon, J.; Khaldy, H.; Reiter, R.J. Melatonin, mitochondria, and cellular bioenergetics. J. Pineal Res 2001, 30, 65–74. [Google Scholar]
  6. Tomas-Zapico, C.; Coto-Montes, A. A proposed mechanism to explain the stimulatory effect of melatonin on antioxidative enzymes. J. Pineal Res 2005, 39, 99–104. [Google Scholar]
  7. Reiter, R.J.; Tan, D.X.; Osuna, C.; Gitto, E. Actions of melatonin in the reduction of oxidative stress. A Review. J. Biomed. Sci 2000, 7, 444–458. [Google Scholar]
  8. De Haan, M.; Wyatt, J.S.; Roth, S.; Vargha-Khadem, F.; Gadian, D.; Mishkin, M. Brain and cognitive-behavioural development after asphyxia at term birth. Dev. Sci 2006, 9, 350–358. [Google Scholar]
  9. Du Plessis, A.J.; Volpe, J.J. Perinatal brain injury in the preterm and term newborn. Curr. Opin. Neurol 2002, 15, 151–157. [Google Scholar]
  10. Hamrick, S.E.; Ferriero, D.M. The injury response in the term newborn brain: Can we neuroprotect? Curr. Opin. Neurol 2003, 16, 147–154. [Google Scholar]
  11. Volpe, J.J. Perinatal brain injury: From pathogenesis to neuroprotection. Ment. Retard. Dev. Disabil. Res. Rev 2001, 7, 56–64. [Google Scholar]
  12. Low, J.A. Determining the contribution of asphyxia to brain damage in the neonate. J. Obstet. Gynaecol. Res 2004, 30, 276–286. [Google Scholar]
  13. Vannucci, S.J.; Hagberg, H. Hypoxia-ischemia in the immature brain. J. Exp. Biol 2004, 207, 3149–3154. [Google Scholar]
  14. Maneru, C.; Junque, C.; Botet, F.; Tallada, M.; Guardia, J. Neuropsychological long-term sequelae of perinatal asphyxia. Brain Inj 2001, 15, 1029–1039. [Google Scholar]
  15. Yager, J.Y.; Armstrong, E.A.; Black, A.M. Treatment of the term newborn with brain injury: Simplicity as the mother of invention. Pediatr. Neurol 2009, 40, 237–243. [Google Scholar]
  16. Vitte, P.A.; Harthe, C.; Lestage, P.; Claustrat, B.; Bobillier, P. Plasma, cerebrospinal fluid, and brain distribution of 14C-melatonin in Rat: A biochemical and autoradiographic study. J. Pineal Res 1988, 5, 437–453. [Google Scholar]
  17. Menendez-Pelaez, A.; Reiter, R.J. Distribution of melatonin in mammalian tissues: The relative importance of nuclear versus cytosolic localization. J. Pineal Res 1993, 15, 59–69. [Google Scholar]
  18. Gupta, Y.K.; Gupta, M.; Kohli, K. Neuroprotective role of melatonin in oxidative stress vulnerable brain. Indian J. Physiol. Pharmacol 2003, 47, 373–386. [Google Scholar]
  19. Rees, S.; Harding, R.; Walker, D. The biological basis of injury and neuroprotection in the fetal and neonatal brain. Int. J. Dev. Neurosci 2011, 29, 551–563. [Google Scholar]
  20. Hilario, E.; Alvarez, A.; Alvarez, F.J.; Gastiasoro, E.; Valls-i-Soler, A. Cellular mechanisms in perinatal hypoxic-ischemic brain injury. Curr. Pediatr. Rev 2006, 2, 131–141. [Google Scholar]
  21. Manev, H.; Uz, T.; Kharlamov, A.; Joo, J.Y. Increased brain damage after stroke or excitotoxic seizures in melatonin-deficient rats. FASEB J 1996, 10, 1546–1551. [Google Scholar]
  22. Li, X.J.; Zhang, L.M.; Gu, J.; Zhang, A.Z.; Sun, F.Y. Melatonin decreases production of hydroxyl radical during cerebral ischemia-reperfusion. Zhongguo Yao Li Xue Bao 1997, 18, 394–396. [Google Scholar]
  23. Cho, S.; Joh, T.H.; Baik, H.H.; Dibinis, C.; Volpe, B.T. Melatonin administration protects CA1 hippocampal neurons after transient forebrain ischemia in rats. Brain Res 1997, 755, 335–338. [Google Scholar]
  24. Kilic, E.; Ozdemir, Y.G.; Bolay, H.; Kelestimur, H.; Dalkara, T. Pinealectomy aggravates and melatonin administration attenuates brain damage in focal ischemia. J. Cereb. Blood Flow MeTable 1999, 19, 511–516. [Google Scholar]
  25. Wakatsuki, A.; Okatani, Y.; Izumiya, C.; Ikenoue, N. Melatonin protects against ischemia and reperfusion-induced oxidative lipid and DNA damage in fetal rat brain. J. Pineal Res 1999, 26, 147–152. [Google Scholar]
  26. Joo, J.Y.; Uz, T.; Manev, H. Opposite effects of pinealectomy and melatonin administration on brain damage following cerebral focal ischemia in rat. Restor. Neurol. Neurosci 1998, 13, 185–191. [Google Scholar]
  27. Cuzzocrea, S.; Costantino, G.; Gitto, E.; Mazzon, E.; Fulia, F.; Serraino, I.; Cordaro, S.; Barberi, I.; De Sarro, A.; Caputi, A.P. Protective effects of melatonin in ischemic brain injury. J. Pineal Res 2000, 29, 217–227. [Google Scholar]
  28. Letechipia-Vallejo, G.; Gonzalez-Burgos, I.; Cervantes, M. Neuroprotective effect of melatonin on brain damage induced by acute global cerebral ischemia in cats. Arch. Med. Res 2001, 32, 186–192. [Google Scholar]
  29. Zhang, J.; Guo, J.D.; Xing, S.H.; Gu, S.L.; Dai, T.J. The protective effects of melatonin on global cerebral ischemia-reperfusion injury in gerbils. Yao Xue Xue Bao 2002, 37, 329–333. [Google Scholar]
  30. Pei, Z.; Pang, S.F.; Cheung, R.T. Administration of melatonin after onset of ischemia reduces the volume of cerebral infarction in a rat middle cerebral artery occlusion stroke model. Stroke 2003, 34, 770–775. [Google Scholar]
  31. Koh, P.O. Melatonin attenuates the focal cerebral ischemic injury by inhibiting the dissociation of pBad from 14-3-3. J. Pineal Res 2008, 44, 101–106. [Google Scholar]
  32. Wang, X.; Figueroa, B.E.; Stavrovskaya, I.G.; Zhang, Y.; Sirianni, A.C.; Zhu, S.; Day, A.L.; Kristal, B.S.; Friedlander, R.M. Methazolamide and melatonin inhibit mitochondrial cytochrome C release and are neuroprotective in experimental models of ischemic injury. Stroke 2009, 40, 1877–1885. [Google Scholar]
  33. Alonso-Alconada, D.; Alvarez, A.; Lacalle, J.; Hilario, E. Histological study of the protective effect of melatonin on neural cells after neonatal hypoxia-ischemia. Histol. Histopathol 2012, 27, 771–783. [Google Scholar]
  34. Carloni, S.; Perrone, S.; Buonocore, G.; Longini, M.; Proietti, F.; Balduini, W. Melatonin protects from the long-term consequences of a neonatal hypoxic-ischemic brain injury in rats. J. Pineal Res 2008, 44, 157–164. [Google Scholar]
  35. Cetinkaya, M.; Alkan, T.; Ozyener, F.; Kafa, I.M.; Kurt, M.A.; Koksal, N. Possible neuroprotective effects of magnesium sulfate and melatonin as both pre- and post-treatment in a neonatal hypoxic-ischemic rat model. Neonatology 2011, 99, 302–310. [Google Scholar]
  36. Ozyener, F.; Cetinkaya, M.; Alkan, T.; Goren, B.; Kafa, I.M.; Kurt, M.A.; Koksal, N. Neuroprotective effects of melatonin administered alone or in combination with topiramate in neonatal hypoxic-ischemic rat model. Restor. Neurol. Neurosci 2012, 30, 435–444. [Google Scholar]
  37. Mattson, M.P.; Guthrie, P.B.; Kater, S.B. Intrinsic factors in the selective vulnerability of hippocampal pyramidal neurons. Prog. Clin. Biol. Res 1989, 317, 333–351. [Google Scholar]
  38. Johnston, M.V. Selective vulnerability in the neonatal brain. Ann. Neurol 1998, 44, 155–156. [Google Scholar]
  39. Northington, F.J.; Ferriero, D.M.; Graham, E.M.; Traystman, R.J.; Martin, L.J. Early neurodegeneration after hypoxia-ischemia in neonatal rat is necrosis while delayed neuronal death is apoptosis. Neurobiol. Dis 2001, 8, 207–219. [Google Scholar]
  40. Hamada, F.; Watanabe, K.; Wakatsuki, A.; Nagai, R.; Shinohara, K.; Hayashi, Y.; Imamura, R.; Fukaya, T. Therapeutic effects of maternal melatonin administration on ischemia/reperfusion-induced oxidative cerebral damage in neonatal rats. Neonatology 2010, 98, 33–40. [Google Scholar]
  41. Watanabe, K.; Hamada, F.; Wakatsuki, A.; Nagai, R.; Shinohara, K.; Hayashi, Y.; Imamura, R.; Fukaya, T. Prophylactic administration of melatonin to the mother throughout pregnancy can protect against oxidative cerebral damage in neonatal rats. J. Matern. Fetal. Neonatal Med 2012, 25, 1254–1259. [Google Scholar]
  42. Takuma, K.; Baba, A.; Matsuda, T. Astrocyte apoptosis: Implications for neuroprotection. Prog. Neurobiol 2004, 72, 111–127. [Google Scholar]
  43. Panickar, K.S.; Norenberg, M.D. Astrocytes in cerebral ischemic injury: Morphological and general considerations. Glia 2005, 50, 287–298. [Google Scholar]
  44. Sizonenko, S.V.; Camm, E.J.; Dayer, A.; Kiss, J.Z. Glial responses to neonatal hypoxic-ischemic injury in the rat cerebral cortex. Int. J. Dev. Neurosci 2008, 26, 37–45. [Google Scholar]
  45. Huang, Z.; Liu, J.; Cheung, P.Y.; Chen, C. Long-term cognitive impairment and myelination deficiency in a rat model of perinatal hypoxic-ischemic brain injury. Brain Res 2009, 1301, 100–109. [Google Scholar]
  46. Xiong, M.; Yang, Y.; Chen, G.Q.; Zhou, W.H. Post-ischemic hypothermia for 24h in P7 rats rescues hippocampal neuron: Association with decreased astrocyte activation and inflammatory cytokine expression. Brain Res. Bull 2009, 79, 351–357. [Google Scholar]
  47. Rothstein, R.P.; Levison, S.W. Gray matter oligodendrocyte progenitors and neurons die caspase-3 mediated deaths subsequent to mild perinatal hypoxic/ischemic insults. Dev. Neurosci 2005, 27, 149–159. [Google Scholar]
  48. Inder, T.E.; Wells, S.J.; Mogridge, N.B.; Spencer, C.; Volpe, J.J. Defining the nature of the cerebral abnormalities in the premature infant: A qualitative magnetic resonance imaging study. J. Pediatr 2003, 143, 171–179. [Google Scholar]
  49. Wang, X.; Hagberg, H.; Zhu, C.; Jacobsson, B.; Mallard, C. Effects of intrauterine inflammation on the developing mouse brain. Brain Res 2007, 1144, 180–185. [Google Scholar]
  50. Olivier, P.; Fontaine, R.H.; Loron, G.; van Steenwinckel, J.; Biran, V.; Massonneau, V.; Kaindl, A.; Dalous, J.; Charriaut-Marlangue, C.; Aigrot, M.S.; et al. Melatonin promotes oligodendroglial maturation of injured white matter in neonatal rats. PLoS One 2009, 4, e7128. [Google Scholar]
  51. Kaur, C.; Sivakumar, V.; Ling, E.A. Melatonin protects periventricular white matter from damage due to hypoxia. J. Pineal Res 2010, 48, 185–193. [Google Scholar]
  52. Villapol, S.; Fau, S.; Renolleau, S.; Biran, V.; Charriaut-Marlangue, C.; Baud, O. Melatonin promotes myelination by decreasing white matter inflammation after neonatal stroke. Pediatr. Res 2011, 69, 51–55. [Google Scholar]
  53. McLean, C.; Ferriero, D. Mechanisms of hypoxic-ischemic injury in the term infant. Semin. Perinatol 2004, 28, 425–432. [Google Scholar]
  54. Sheldon, R.A.; Jiang, X.; Francisco, C.; Christen, S.; Vexler, Z.S.; Tauber, M.G.; Ferriero, D.M. Manipulation of antioxidant pathways in neonatal murine brain. Pediatr. Res 2004, 56, 656–662. [Google Scholar]
  55. McQuillen, P.S.; Ferriero, D.M. Selective vulnerability in the developing central nervous system. Pediatr. Neurol 2004, 30, 227–235. [Google Scholar]
  56. Miller, S.L.; Yan, E.B.; Castillo-Melendez, M.; Jenkin, G.; Walker, D.W. Melatonin provides neuroprotection in the late-gestation fetal sheep brain in response to umbilical cord occlusion. Dev. Neurosci 2005, 27, 200–210. [Google Scholar]
  57. Tutunculer, F.; Eskiocak, S.; Basaran, U.N.; Ekuklu, G.; Ayvaz, S.; Vatansever, U. The protective role of melatonin in experimental hypoxic brain damage. Pediatr. Int 2005, 47, 434–439. [Google Scholar]
  58. Fulia, F.; Gitto, E.; Cuzzocrea, S.; Reiter, R.J.; Dugo, L.; Gitto, P.; Barberi, S.; Cordaro, S.; Barberi, I. Increased levels of malondialdehyde and nitrite/nitrate in the blood of asphyxiated newborns: Reduction by melatonin. J. Pineal Res 2001, 31, 343–349. [Google Scholar]
  59. Kohen, R.; Nyska, A. Oxidation of biological systems: Oxidative stress phenomena, antioxidants, redox reactions, and methods for their quantification. Toxicol. Pathol 2002, 30, 620–650. [Google Scholar]
  60. Arneson, K.O.; Roberts, L.J., 2nd. Measurement of products of docosahexaenoic acid peroxidation, neuroprostanes, and neurofurans. Methods Enzymol 2007, 433, 127–143. [Google Scholar]
  61. Song, W.L.; Lawson, J.A.; Reilly, D.; Rokach, J.; Chang, C.T.; Giasson, B.; FitzGerald, G.A. Neurofurans, novel indices of oxidant stress derived from docosahexaenoic acid. J. Biol. Chem 2008, 283, 6–16. [Google Scholar]
  62. Signorini, C.; Ciccoli, L.; Leoncini, S.; Carloni, S.; Perrone, S.; Comporti, M.; Balduini, W.; Buonocore, G. Free iron, total F-isoprostanes and total F-neuroprostanes in a model of neonatal hypoxic-ischemic encephalopathy: Neuroprotective effect of melatonin. J. Pineal Res 2009, 46, 148–154. [Google Scholar]
  63. Balduini, W.; Carloni, S.; Perrone, S.; Bertrando, S.; Tataranno, M.L.; Negro, S.; Proietti, F.; Longini, M.; Buonocore, G. The use of melatonin in hypoxic-ischemic brain damage: An experimental study. J. Matern. Fetal. Neonatal Med 2012, 25, 119–124. [Google Scholar]
  64. Welin, A.K.; Svedin, P.; Lapatto, R.; Sultan, B.; Hagberg, H.; Gressens, P.; Kjellmer, I.; Mallard, C. Melatonin reduces inflammation and cell death in white matter in the mid-gestation fetal sheep following umbilical cord occlusion. Pediatr. Res 2007, 61, 153–158. [Google Scholar]
  65. Eskiocak, S.; Tutunculer, F.; Basaran, U.N.; Taskiran, A.; Cakir, E. The Effect of melatonin on protein oxidation and nitric oxide in the brain tissue of hypoxic neonatal rats. Brain Dev 2007, 29, 19–24. [Google Scholar]
  66. Witko-Sarsat, V.; Friedlander, M.; Capeillere-Blandin, C.; Nguyen-Khoa, T.; Nguyen, A.T.; Zingraff, J.; Jungers, P.; Descamps-Latscha, B. Advanced oxidation protein products as a novel marker of oxidative stress in uremia. Kidney Int 1996, 49, 1304–1313. [Google Scholar]
  67. Andrabi, S.A.; Sayeed, I.; Siemen, D.; Wolf, G.; Horn, T.F. Direct Inhibition of the mitochondrial permeability transition pore: A possible mechanism responsible for anti-apoptotic effects of melatonin. FASEB J 2004, 18, 869–871. [Google Scholar]
  68. Wang, X.; Zhu, S.; Pei, Z.; Drozda, M.; Stavrovskaya, I.G.; Del Signore, S.J.; Cormier, K.; Shimony, E.M.; Wang, H.; Ferrante, R.J.; et al. Inhibitors of cytochrome c release with therapeutic potential for huntington’s disease. J. Neurosci 2008, 28, 9473–9485. [Google Scholar]
  69. Jou, M.J.; Peng, T.I.; Reiter, R.J.; Jou, S.B.; Wu, H.Y.; Wen, S.T. Visualization of the antioxidative effects of melatonin at the mitochondrial level during oxidative stress-induced apoptosis of rat brain astrocytes. J. Pineal Res 2004, 37, 55–70. [Google Scholar]
  70. Kilic, E.; Kilic, U.; Reiter, R.J.; Bassetti, C.L.; Hermann, D.M. Tissue-plasminogen activator-induced ischemic brain injury is reversed by melatonin: Role of inos and akt. J. Pineal Res 2005, 39, 151–155. [Google Scholar]
  71. Jang, M.H.; Jung, S.B.; Lee, M.H.; Kim, C.J.; Oh, Y.T.; Kang, I.; Kim, J.; Kim, E.H. Melatonin attenuates amyloid beta25–35-induced apoptosis in mouse microglial bv2 cells. Neurosci. Lett 2005, 380, 26–31. [Google Scholar]
  72. Ebadi, M.; Sharma, S.K.; Ghafourifar, P.; Brown-Borg, H.; El Refaey, H. Peroxynitrite in the pathogenesis of parkinson’s disease and the neuroprotective role of metallothioneins. Methods Enzymol 2005, 396, 276–298. [Google Scholar]
  73. Alvira, D.; Tajes, M.; Verdaguer, E.; Acuna-Castroviejo, D.; Folch, J.; Camins, A.; Pallas, M. Inhibition of the cdk5/p25 fragment formation may explain the antiapoptotic effects of melatonin in an experimental model of parkinson’s disease. J. Pineal Res 2006, 40, 251–258. [Google Scholar]
  74. Ling, X.; Zhang, L.M.; Lu, S.D.; Li, X.J.; Sun, F.Y. Protective effect of melatonin on injuried cerebral neurons is associated with bcl-2 protein over-expression. Zhongguo Yao Li Xue Bao 1999, 20, 409–414. [Google Scholar]
  75. Sun, F.Y.; Lin, X.; Mao, L.Z.; Ge, W.H.; Zhang, L.M.; Huang, Y.L.; Gu, J. Neuroprotection by melatonin against ischemic neuronal injury associated with modulation of DNA damage and repair in the rat following a transient cerebral ischemia. J. Pineal Res 2002, 33, 48–56. [Google Scholar]
  76. Chung, S.Y.; Han, S.H. Melatonin attenuates kainic acid-induced hippocampal neurodegeneration and oxidative stress through microglial inhibition. J. Pineal Res 2003, 34, 95–102. [Google Scholar]
  77. Feng, Z.; Cheng, Y.; Zhang, J.T. Long-term effects of melatonin or 17 beta-estradiol on improving spatial memory performance in cognitively impaired, ovariectomized adult rats. J. Pineal Res 2004, 37, 198–206. [Google Scholar]
  78. Deng, Y.Q.; Xu, G.G.; Duan, P.; Zhang, Q.; Wang, J.Z. Effects of melatonin on wortmannin-induced TAU hyperphosphorylation. Acta Pharmacol. Sin 2005, 26, 519–526. [Google Scholar]
  79. Rosenstein, R.E.; Cardinali, D.P. Central gabaergic mechanisms as targets for melatonin activity in brain. Neurochem. Int 1990, 17, 373–379. [Google Scholar]
  80. Molina-Carballo, A.; Munoz-Hoyos, A.; Sanchez-Forte, M.; Uberos-Fernandez, J.; Moreno-Madrid, F.; Acuna-Castroviejo, D. Melatonin increases following convulsive seizures may be related to its anticonvulsant properties at physiological concentrations. Neuropediatrics 2007, 38, 122–125. [Google Scholar]
  81. Prada, C.; Udin, S.B.; Wiechmann, A.F.; Zhdanova, I.V. Stimulation of melatonin receptors decreases calcium levels in xenopus tectal cells by activating gaba(c) receptors. J. Neurophysiol 2005, 94, 968–978. [Google Scholar]
  82. Prada, C.; Udin, S.B. Melatonin decreases calcium levels in retinotectal axons of xenopus laevis by indirect activation of group iii metabotropic glutamate receptors. Brain Res 2005, 1053, 67–76. [Google Scholar]
  83. Buonocore, G.; Perrone, S.; Bracci, R. Free radicals and brain damage in the newborn. Biol. Neonate 2001, 79, 180–186. [Google Scholar]
  84. Blomgren, K.; Hagberg, H. Free radicals, mitochondria, and hypoxia-ischemia in the developing brain. Free Radic. Biol. Med 2006, 40, 388–397. [Google Scholar]
  85. Kumar, A.; Mittal, R.; Khanna, H.D.; Basu, S. Free radical injury and blood-brain barrier permeability in hypoxic-ischemic encephalopathy. Pediatrics 2008, 122, 722–727. [Google Scholar]
  86. Alonso-Alconada, D.; Hilario, E.; Alvarez, F.J.; Alvarez, A. Apoptotic cell death correlates with ros overproduction and early cytokine expression after hypoxia-ischemia in fetal lambs. Reprod. Sci 2012, 19, 754–763. [Google Scholar]
  87. Watanabe, K.; Wakatsuki, A.; Shinohara, K.; Ikenoue, N.; Yokota, K.; Fukaya, T. Maternally administered melatonin protects against ischemia and reperfusion-induced oxidative mitochondrial damage in premature fetal rat brain. J. Pineal Res 2004, 37, 276–280. [Google Scholar]
  88. Hutton, L.C.; Abbass, M.; Dickinson, H.; Ireland, Z.; Walker, D.W. Neuroprotective properties of melatonin in a model of birth asphyxia in the spiny mouse (Acomys cahirinus). Dev. Neurosci 2009, 31, 437–451. [Google Scholar]
  89. Fu, J.; Zhao, S.D.; Liu, H.J.; Yuan, Q.H.; Liu, S.M.; Zhang, Y.M.; Ling, E.A.; Hao, A.J. Melatonin promotes proliferation and differentiation of neural stem cells subjected to hypoxia in vitro. J. Pineal Res 2011, 51, 104–112. [Google Scholar]
  90. Kilic, U.; Kilic, E.; Reiter, R.J.; Bassetti, C.L.; Hermann, D.M. Signal transduction pathways involved in melatonin-induced neuroprotection after focal cerebral ischemia in mice. J. Pineal Res 2005, 38, 67–71. [Google Scholar]
  91. Koh, P.O. Melatonin prevents the injury-induced decline of akt/forkhead transcription factors phosphorylation. J. Pineal Res 2008, 45, 199–203. [Google Scholar]
  92. Zhou, J.; Zhang, S.; Zhao, X.; Wei, T. Melatonin impairs nadph oxidase assembly and decreases superoxide anion production in microglia exposed to amyloid-beta1–42. J. Pineal Res 2008, 45, 157–165. [Google Scholar]
  93. Koh, P.O. Melatonin prevents ischemic brain injury through activation of the mtor/p70s6 kinase signaling pathway. Neurosci. Lett 2008, 444, 74–78. [Google Scholar]
  94. Fjaerli, O.; Lund, T.; Osterud, B. The effect of melatonin on cellular activation processes in human blood. J. Pineal Res 1999, 26, 50–55. [Google Scholar]
  95. Baykal, A.; Iskit, A.B.; Hamaloglu, E.; Guc, M.O.; Hascelik, G.; Sayek, I. Melatonin modulates mesenteric blood flow and TNFα concentrations after lipopolysaccharide challenge. Eur. J. Surg 2000, 166, 722–727. [Google Scholar]
  96. Silva, S.O.; Rodrigues, M.R.; Ximenes, V.F.; Bueno-da-Silva, A.E.; Amarante-Mendes, G.P.; Campa, A. Neutrophils as a specific target for melatonin and kynuramines: Effects on cytokine release. J. Neuroimmunol 2004, 156, 146–152. [Google Scholar]
  97. Wang, H.; Wei, W.; Shen, Y.X.; Dong, C.; Zhang, L.L.; Wang, N.P.; Yue, L.; Xu, S.Y. Protective effect of melatonin against liver injury in mice induced by bacillus calmette-guerin plus lipopolysaccharide. World J. Gastroenterol 2004, 10, 2690–2696. [Google Scholar]
  98. Perianayagam, M.C.; Oxenkrug, G.F.; Jaber, B.L. Immune-modulating effects of melatonin, N-acetylserotonin, and N-acetyldopamine. Ann. N.Y. Acad. Sci 2005, 1053, 386–393. [Google Scholar]
  99. Carrillo-Vico, A.; Lardone, P.J.; Fernandez-Santos, J.M.; Martin-Lacave, I.; Calvo, J.R.; Karasek, M.; Guerrero, J.M. Human lymphocyte-synthesized melatonin is involved in the regulation of the interleukin-2/interleukin-2 receptor system. J. Clin. Endocrinol. MeTable 2005, 90, 992–1000. [Google Scholar]
  100. Gitto, E.; Reiter, R.J.; Sabatino, G.; Buonocore, G.; Romeo, C.; Gitto, P.; Bugge, C.; Trimarchi, G.; Barberi, I. Correlation among cytokines, bronchopulmonary dysplasia and modality of ventilation in preterm newborns: Improvement with melatonin treatment. J. Pineal Res 2005, 39, 287–293. [Google Scholar]
  101. Steinhilber, D.; Brungs, M.; Werz, O.; Wiesenberg, I.; Danielsson, C.; Kahlen, J.P.; Nayeri, S.; Schrader, M.; Carlberg, C. The nuclear receptor for melatonin represses 5-lipoxygenase gene expression in human B lymphocytes. J. Biol. Chem 1995, 270, 7037–7040. [Google Scholar]
  102. Mayo, J.C.; Sainz, R.M.; Tan, D.X.; Hardeland, R.; Leon, J.; Rodriguez, C.; Reiter, R.J. Anti-inflammatory actions of melatonin and its metabolites, N1-acetyl-N2-formyl-5- methoxykynuramine (AFMK) and N1-acetyl-5-methoxykynuramine (AMK), in macrophages. J. Neuroimmunol 2005, 165, 139–149. [Google Scholar]
  103. Deng, W.G.; Tang, S.T.; Tseng, H.P.; Wu, K.K. Melatonin suppresses macrophage cyclooxygenase-2 and inducible nitric oxide synthase expression by inhibiting p52 acetylation and binding. Blood 2006, 108, 518–524. [Google Scholar]
  104. Cardinali, D.P.; Ritta, M.N.; Fuentes, A.M.; Gimeno, M.F.; Gimeno, A.L. Prostaglandin E release by rat medial basal hypothalamus in vitro. Inhibition by melatonin at submicromolar concentrations. Eur. J. Pharmacol 1980, 67, 151–153. [Google Scholar]
  105. Cardinali, D.P.; Ritta, M.N. The role of prostaglandins in neuroendocrine junctions: Studies in the pineal gland and the hypothalamus. Neuroendocrinology 1983, 36, 152–160. [Google Scholar]
  106. Carrillo-Vico, A.; Garcia-Maurino, S.; Calvo, J.R.; Guerrero, J.M. Melatonin counteracts the inhibitory effect of PGE2 on IL-2 production in human lymphocytes via its mt1 membrane receptor. FASEB J 2003, 17, 755–757. [Google Scholar]
  107. Bilici, D.; Akpinar, E.; Kiziltunc, A. Protective effect of melatonin in carrageenan-induced acute local inflammation. Pharmacol. Res 2002, 46, 133–139. [Google Scholar]
  108. Acuna-Castroviejo, D.; Escames, G.; Lopez, L.C.; Hitos, A.B.; Leon, J. Melatonin and nitric oxide: Two required antagonists for mitochondrial homeostasis. Endocrine 2005, 27, 159–168. [Google Scholar]
  109. Leon, J.; Macias, M.; Escames, G.; Camacho, E.; Khaldy, H.; Martin, M.; Espinosa, A.; Gallo, M.A.; Acuna-Castroviejo, D. Structure-related inhibition of calmodulin-dependent neuronal nitric-oxide synthase activity by melatonin and synthetic kynurenines. Mol. Pharmacol 2000, 58, 967–975. [Google Scholar]
  110. Leon, J.; Escames, G.; Rodriguez, M.I.; Lopez, L.C.; Tapias, V.; Entrena, A.; Camacho, E.; Carrion, M.D.; Gallo, M.A.; Espinosa, A.; et al. Inhibition of neuronal nitric oxide synthase activity by N1-acetyl-5-methoxykynuramine, a brain metabolite of melatonin. J. Neurochem 2006, 98, 2023–2033. [Google Scholar]
  111. Jimenez-Ortega, V.; Cano, P.; Cardinali, D.P.; Esquifino, A.I. 24-Hour variation in gene expression of redox pathway enzymes in rat hypothalamus: Effect of melatonin treatment. Redox Rep 2009, 14, 132–138. [Google Scholar]
  112. Tapias, V.; Escames, G.; Lopez, L.C.; Lopez, A.; Camacho, E.; Carrion, M.D.; Entrena, A.; Gallo, M.A.; Espinosa, A.; Acuna-Castroviejo, D. Melatonin and its brain metabolite N(1)-acetyl-5-methoxykynuramine prevent mitochondrial nitric oxide synthase induction in parkinsonian mice. J. Neurosci. Res 2009, 87, 3002–3010. [Google Scholar]
  113. Leon, J.; Vives, F.; Crespo, E.; Camacho, E.; Espinosa, A.; Gallo, M.A.; Escames, G.; Acuna-Castroviejo, D. Modification of nitric oxide synthase activity and neuronal response in rat striatum by melatonin and kynurenine derivatives. J. Neuroendocrinol 1998, 10, 297–302. [Google Scholar]
  114. Chandrasekaran, A.; Ponnambalam, G.; Kaur, C. Domoic acid-induced neurotoxicity in the hippocampus of adult rats. Neurotox Res 2004, 6, 105–117. [Google Scholar]
  115. Escames, G.; Khaldy, H.; Leon, J.; Gonzalez, L.; Acuna-Castroviejo, D. Changes in iNOS activity, oxidative stress and melatonin levels in hypertensive patients treated with lacidipine. J. Hypertens 2004, 22, 629–635. [Google Scholar]
  116. Escames, G.; Acuna-Castroviejo, D.; Lopez, L.C.; Tan, D.X.; Maldonado, M.D.; Sanchez-Hidalgo, M.; Leon, J.; Reiter, R.J. Pharmacological utility of melatonin in the treatment of septic shock: Experimental and clinical evidence. J. Pharm. Pharmacol 2006, 58, 1153–1165. [Google Scholar]
  117. Escames, G.; Lopez, L.C.; Tapias, V.; Utrilla, P.; Reiter, R.J.; Hitos, A.B.; Leon, J.; Rodriguez, M.I.; Acuna-Castroviejo, D. Melatonin counteracts inducible mitochondrial nitric oxide synthase-dependent mitochondrial dysfunction in skeletal muscle of septic mice. J. Pineal Res 2006, 40, 71–78. [Google Scholar]
  118. Lopez, L.C.; Escames, G.; Tapias, V.; Utrilla, P.; Leon, J.; Acuna-Castroviejo, D. Identification of an inducible nitric oxide synthase in diaphragm mitochondria from septic mice: Its relation with mitochondrial dysfunction and prevention by melatonin. Int. J. Biochem. Cell Biol 2006, 38, 267–278. [Google Scholar]
  119. Srinivasan, V.; Pandi-Perumal, S.R.; Spence, D.W.; Kato, H.; Cardinali, D.P. Melatonin in septic shock: Some recent concepts. J. Crit. Care 2010, 25, 656, e1–656.e6.. [Google Scholar]
  120. Lopez, L.C.; Escames, G.; Ortiz, F.; Ros, E.; Acuna-Castroviejo, D. Melatonin restores the mitochondrial production of ATP in septic mice. Neuro Endocrinol. Lett 2006, 27, 623–630. [Google Scholar]
  121. Escames, G.; Lopez, L.C.; Ortiz, F.; Lopez, A.; Garcia, J.A.; Ros, E.; Acuna-Castroviejo, D. Attenuation of cardiac mitochondrial dysfunction by melatonin in septic mice. FEBS J 2007, 274, 2135–2147. [Google Scholar]
  122. Pei, Z.; Cheung, R.T. Pretreatment with melatonin exerts anti-inflammatory effects against ischemia/reperfusion injury in a rat middle cerebral artery occlusion stroke model. J. Pineal Res 2004, 37, 85–91. [Google Scholar]
  123. Lee, M.Y.; Kuan, Y.H.; Chen, H.Y.; Chen, T.Y.; Chen, S.T.; Huang, C.C.; Yang, I.P.; Hsu, Y.S.; Wu, T.S.; Lee, E.J. Intravenous administration of melatonin reduces the intracerebral cellular inflammatory response following transient focal cerebral ischemia in rats. J. Pineal Res 2007, 42, 297–309. [Google Scholar]
  124. Koh, P.O. Melatonin regulates nitric oxide synthase expression in ischemic brain injury. J. Vet. Med. Sci 2008, 70, 747–750. [Google Scholar]
  125. Mohan, N.; Sadeghi, K.; Reiter, R.J.; Meltz, M.L. The neurohormone melatonin inhibits cytokine, mitogen and ionizing radiation induced NF-κB. Biochem. Mol. Biol. Int 1995, 37, 1063–1070. [Google Scholar]
  126. Reiter, R.J.; Calvo, J.R.; Karbownik, M.; Qi, W.; Tan, D.X. Melatonin and its relation to the immune system and inflammation. Ann. N.Y. Acad. Sci 2000, 917, 376–386. [Google Scholar]
  127. Hardeland, R. Antioxidative protection by melatonin: Multiplicity of mechanisms from radical detoxification to radical avoidance. Endocrine 2005, 27, 119–130. [Google Scholar]
  128. Kaur, C.; Sivakumar, V.; Lu, J.; Tang, F.R.; Ling, E.A. Melatonin attenuates hypoxia-induced ultrastructural changes and increased vascular permeability in the developing hippocampus. Brain Pathol 2008, 18, 533–547. [Google Scholar]
  129. Jan, J.E.; Wasdell, M.B.; Freeman, R.D.; Bax, M. Evidence supporting the use of melatonin in short gestation infants. J. Pineal Res 2007, 42, 22–27. [Google Scholar]
  130. Gitto, E.; Reiter, R.J.; Cordaro, S.P.; La Rosa, M.; Chiurazzi, P.; Trimarchi, G.; Gitto, P.; Calabro, M.P.; Barberi, I. Oxidative and inflammatory parameters in respiratory distress syndrome of preterm newborns: Beneficial effects of melatonin. Am. J. Perinatol 2004, 21, 209–216. [Google Scholar]
  131. Gitto, E.; Reiter, R.J.; Amodio, A.; Romeo, C.; Cuzzocrea, E.; Sabatino, G.; Buonocore, G.; Cordaro, V.; Trimarchi, G.; Barberi, I. Early indicators of chronic lung disease in preterm infants with respiratory distress syndrome and their inhibition by melatonin. J. Pineal Res 2004, 36, 250–255. [Google Scholar]
  132. Buonocore, G.; Groenendaal, F. Anti-oxidant strategies. Semin. Fetal. Neonatal Med 2007, 12, 287–295. [Google Scholar]
  133. Robertson, N.J.; Faulkner, S.; Fleiss, B.; Bainbridge, A.; Andorka, C.; Price, D.; Powell, E.; Lecky-Thompson, L.; Thei, L.; Chandrasekaran, M.; et al. Melatonin augments hypothermic neuroprotection in a perinatal asphyxia model. Brain 2013, 136, 90–105. [Google Scholar]
  134. Waldhauser, F.; Waldhauser, M.; Lieberman, H.R.; Deng, M.H.; Lynch, H.J.; Wurtman, R.J. Bioavailability of oral melatonin in humans. Neuroendocrinology 1984, 39, 307–313. [Google Scholar]
  135. Aldhous, M.; Franey, C.; Wright, J.; Arendt, J. Plasma concentrations of melatonin in man following oral absorption of different preparations. Br. J. Clin. Pharmacol 1985, 19, 517–521. [Google Scholar]
  136. Lane, E.A.; Moss, H.B. Pharmacokinetics of melatonin in man: First pass hepatic metabolism. J. Clin. Endocrinol. MeTable 1985, 61, 1214–1216. [Google Scholar]
  137. Merchant, N.M.; Azzopardi, D.V.; Hawwa, A.F.; McElnay, J.C.; Middleton, B.; Arendt, J.; Arichi, T.; Gressens, P.; Edwards, A.D. Pharmacokinetics of Melatonin in Preterm Infants. Br. J. Clin. Pharmacol. 2013. [Google Scholar] [CrossRef]
Figure 1. Nissl-stained (AC), myelin basic protein (DF) and glial fibrillary acidic protein (GI) immunolabeled brain sections corresponding to the surrounding areas of the CA1 region of the hippocampus and the external capsule showing cell loss (B), myelination deficit (E) and reactive gliosis (H) after hypoxia-ischemia and recovery after melatonin administration. Seven-day old rats were subjected to hypoxia-ischemia (left common carotid artery ligated and then 8% oxygen for 2 h) and sacrificed seven days after the injury. Pups without ischemia or hypoxia served as controls (Sham group). Bar: 100 μm.
Figure 1. Nissl-stained (AC), myelin basic protein (DF) and glial fibrillary acidic protein (GI) immunolabeled brain sections corresponding to the surrounding areas of the CA1 region of the hippocampus and the external capsule showing cell loss (B), myelination deficit (E) and reactive gliosis (H) after hypoxia-ischemia and recovery after melatonin administration. Seven-day old rats were subjected to hypoxia-ischemia (left common carotid artery ligated and then 8% oxygen for 2 h) and sacrificed seven days after the injury. Pups without ischemia or hypoxia served as controls (Sham group). Bar: 100 μm.
Ijms 14 09379f1
Table. Summary of the experimental evidence regarding the beneficial effects of melatonin.
Table. Summary of the experimental evidence regarding the beneficial effects of melatonin.
TargetEffectReferences
Brain Protection
Infarct volume [3436]
Sensorimotor asymmetry [34]
Learning deficits [34]
Morphologically well preserved neurons [33,40,41]
GFAP expression [33]
MBP expression [33,5052]
Antioxidant
Lipid peroxidation and MDA production [5658]
Iso- and neuroprostanes and neurofurans [6264]
Protein oxidation [65]
Catalase’s activity [57]
Hydroxyl formation [56]
Nitrite/nitrate levels [58]
Anti-apoptotic
Cytochrome c release [32,67,68]
Caspase-1 and Caspase-3 activation [32,67,6973,88,89]
Bcl-xL and Bcl-2 expression [70,71,74,75,89]
Bax expression [71]
Poly-ADP-ribose-polymerase cleavage [72]
Mitochondrial transition pore opening [67,69]
TUNEL-positive cells/DNA breaks [3133,35,36,64,72,7578]
Cytosolic calcium concentrations [81,82]
Oxidative mitochondria damage [87]
Mitochondrial respiratory activity [40,41]
Oxidative stress [40,41]
Fractin levels [88]
Bcl-2/Bax ratio [89]
MAP kinase, JNK1/2 and ERK 1/2 [31,72,90]
Bad dephosphorylation [31]
Anti-inflammatory
Interleukin-6, Interleukin-8 and Tumor Necrosis Factor- α [94100,102,122,123,125,126]
5-lipoxygenase and Cyclooxyenase-2 [101103,127]
Prostaglandin [104,106]
NO, nNOS, iNOS and VEGF [107112,124,128]
Macrophage infiltration [88]
ED1 positive cells [63]

Share and Cite

MDPI and ACS Style

Alonso-Alconada, D.; Álvarez, A.; Arteaga, O.; Martínez-Ibargüen, A.; Hilario, E. Neuroprotective Effect of Melatonin: A Novel Therapy against Perinatal Hypoxia-Ischemia. Int. J. Mol. Sci. 2013, 14, 9379-9395. https://doi.org/10.3390/ijms14059379

AMA Style

Alonso-Alconada D, Álvarez A, Arteaga O, Martínez-Ibargüen A, Hilario E. Neuroprotective Effect of Melatonin: A Novel Therapy against Perinatal Hypoxia-Ischemia. International Journal of Molecular Sciences. 2013; 14(5):9379-9395. https://doi.org/10.3390/ijms14059379

Chicago/Turabian Style

Alonso-Alconada, Daniel, Antonia Álvarez, Olatz Arteaga, Agustín Martínez-Ibargüen, and Enrique Hilario. 2013. "Neuroprotective Effect of Melatonin: A Novel Therapy against Perinatal Hypoxia-Ischemia" International Journal of Molecular Sciences 14, no. 5: 9379-9395. https://doi.org/10.3390/ijms14059379

Article Metrics

Back to TopTop