Abstract
The excessive intake of alcohol is a serious public health problem, especially given the severe damage provoked by chronic or prenatal exposure to alcohol that affects many physiological processes, such as memory, motor function, and cognitive abilities. This damage is related to the ethanol oxidation in the brain. The metabolism of ethanol to acetaldehyde and then to acetate is associated with the production of reactive oxygen species that accentuate the oxidative state of cells. This metabolism of ethanol can induce the oxidation of the fatty acids in phospholipids, and the bioactive aldehydes produced are known to be associated with neurotoxicity and neurodegeneration. As such, here we will review the role of lipids in the neuronal damage induced by ethanol-related oxidative stress and the role that lipids play in the related compensatory or defense mechanisms.
1. Introduction
A clear relationship has been established between ethanol intake, addiction and dependency [1–3], and several risk factors for chronic disease and injury [4]. Indeed, the public health problem associated with increased alcohol consumption and alcoholism [5, 6] is becoming ever more severe due to the increased economic burden of the complications on the health national systems and the cost of the relevant treatments [4, 7–10]. Alcoholism provokes high rates of mortality and it increases in risk of several disabling disorders [4, 11]. Such damage can be classified in function of the organs involved (liver, kidney, heart, brain, etc.), the type of intake (acute or chronic), or the subject's age at the time of exposure to ethanol (prenatal, neonatal, or adult). In summary, ethanol has several negative health effects, especially if we consider prenatal exposure where the brain is a major target for the damage provoked.
2. Effects of Ethanol in the Brain
Ethanol has many effects in the brain depending on the age of exposure (prenatal, postnatal, or adult). For example, aggressive behavior and depression are observed after acute postnatal exposure to ethanol, possibly due to a decrease in circulating tryptophan, followed by the depletion of serotonin in the brain [12, 13]. Another effect of acute postnatal alcohol exposure is related to impaired impulsive and control behavior [14, 15], although few in vivo studies have focused on this issue. Cognitive performance has been associated with specific prefrontal cortical regions in Rhesus Macaque monkeys [11] and GABA receptors in this structure have been implicated in the effects of acute postnatal ethanol exposure [16–18]. Indeed, GABA was found to be a mediator in ethanol-induced ataxia [18, 19].
The most severe alcohol-related damage is found following acute prenatal or chronic pre- and postnatal ethanol exposure, effects that have been associated with a loss of neurons (Table 1). In terms of prenatal exposure, the babies born to women that drink alcohol excessively during pregnancy may suffer from fetal alcohol syndrome [20], a condition characterized by specific craniofacial abnormalities, pre- and postnatal growth deficiencies, and nervous system dysfunction that is manifested as persistent intellectual, behavioral, and neurological defects [5, 21]. These latter symptoms have been related to neurodegeneration in experimental animal models (see Table 1 for a summary of some of the available literature).
Table 1.
Age of exposure | Species (model) | Frequency of administration | Structural CNS changes | Behavioral, intellectual, or other effects | References |
---|---|---|---|---|---|
Prenatal | Rat | Chronic | Reduced number of neurons and dendritic spines in the hippocampus and pyramidal tracts | Memory, spatial learning | [22] |
The cerebellum is most sensitive to alcohol-induced Purkinje cell loss | Cerebellar disorders (ataxia, cognitive, behavioral, and affective disturbances) | [23, 24] | |||
Cholinergic neurons loss | Spontaneous alternation, spatial working memory | [25, 26] | |||
Human | Chronic | Reduction in gray and white matter in the hippocampus, amygdala, thalamus, caudate, putamen, and globus pallidus | Cognitive, behavioral, and neurological impairments | [27, 28] | |
| |||||
Young | Human | Chronic | Reduced white matter, corpus callosum, and hippocampal volumes | Impairment in neurocognitive tests, including those measuring memory, attention, visuospatial skills, and executive function | [29, 30] |
Reduced oxygen consumption in the subcallosal, anterior cingulate, left prefrontal, and bilateral insular regions | Dysfunction during spatial working memory and simple motor tasks | [31, 32] | |||
| |||||
Adult | Human | Chronic | Reduced volume in the diencephalon, cerebral cortex, hippocampus, and white matter | Progressive cognitive dysfunction and loss of neural plasticity due to reduced GABAergic inhibition and increased glutamatergic excitation | [33] |
Rat | Chronic | Corpus callosum ultrastructure | Cognitive and motor function | [34] | |
Human | Chronic | Frontal and temporal lobes | Attention, impulsivity, verbal memory, and impaired cognition | [35, 36] | |
Human or monkey | Chronic | Hypothalamus D3 and 5HT1A neuronal receptors | Alcohol dependency | [37, 38] | |
Human | Chronic | Nucleus basalis Meynert | Loss cognitive disorders and dementia | [39–42] | |
Cerebellar atrophy, Purkinje cell loss | Wernicke's encephalopathy, cognitive and emotional dysfunction | [43, 44] | |||
Peripheral nerves stimulation | Withdrawal-induced hyperalgesia | [45] | |||
Hemorrhage in the ventral diencephalon, mesencephalon, and Basal ganglia, and severe white matter edema in the cerebral hemispheres and pontine nuclei and medullary tegmental | Cognitive impairment, necrosis, and death | [46, 47] |
Chronic alcohol exposure has been associated with permanent neuronal loss in brain regions like the hippocampus and cerebellum. Moreover, in vivo studies have demonstrated neurological effects following chronic ethanol exposure in young and adult populations, with deterioration in memory, motor function, cognition, and so forth. All these effects could be due to neurotoxicity or neurodegeneration, and there is evidence that oxidative stress associated with ethanol metabolism is involved.
3. The Pharmacokinetics of Ethanol
The ethanol concentration that can be found in blood following its ingestion depends on its pharmacokinetics (PK). PK determines not only the time-course and persistence of ethanol in blood but also the amount of alcohol and its metabolic products that accumulate in different tissues, and hence their pharmacological and toxicological responses [48].
3.1. Absorption
In adults, the ethanol ingested is almost completely and instantly absorbed by passive diffusion, reaching a peak concentration in humans between 30 and 90 min. Absorption is more efficient in the small intestine than in stomach [49], a difference in absorption that is due to two factors. First, the thickness of mucus that protects the stomach appears to have a resistance ~16 times greater than that which protects the small intestine [49, 50], which also has a greater intestinal absorption surface due to the presence villi and microvilli [51]. The second difference reflects the speed of stirring caused by peristalsis, which is more important in the small intestine than in the stomach, playing a role in gastric emptying and in the intestinal transit time [49].
In addition, the presence of food is another factor that modifies the absorption rate [52], mainly as food reduces gastric emptying and ethanol is absorbed more slowly [53]. Solid food intake can reduce the ethanol absorption rate by 30% and it has been suggested that this effect is due to the need for food digestion prior to absorption process. As such, if food is taken in as a liquid then it would not produce this effect [49, 54]. Moreover, a small amount of ethanol can be oxidized to acetaldehyde by alcohol dehydrogenase (ADH) classes I and IV [52, 55] in the stomach and intestine. This acetaldehyde can be absorbed along with ethanol and metabolized by the liver or other tissues.
3.2. First Pass Metabolism and the Distribution of Ethanol
The amount of alcohol in any given tissue depends on its relative concentration in the blood, which is a function of first pass metabolism [49], that is, the oxidation of ethanol in the stomach, intestine, and liver.
Most first pass metabolism occurs in the liver [49, 55] and the rate-limiting step is the oxidation of ethanol to acetaldehyde. This reaction is catalyzed by proteins of the ADH family [56], of which class I (ADH1) and III (ADH3) enzymes metabolize ethanol in the liver [57, 58]. These two types of enzymes differ in their Km, with ADH1 having a low Km while ADH3 has a high Km value [57, 59]. Consequently ADH3 plays a more important role in the metabolism of alcohol at high concentrations. In addition, microsomal ethanol oxidizing system (MEOS) and catalase contribute to the metabolism of alcohol in specific circumstances, such as high ethanol concentrations [48, 60].
The acetaldehyde produced by the oxidation of ethanol is thereafter transformed to acetate by aldehyde dehydrogenase (ALDH) [61], which can be further metabolized through the tricarboxylic acid cycle to generate energy, or these metabolites can be deposited in the plasma [62, 63]. Indeed, increases in acetate but not acetaldehyde can be detected in human plasma after ethanol intake [64, 65] (Figure 1).
The efficiency of ethanol metabolism is dependent on the enzymatic activity and pathways involved. It has been reported that ADH, cytochrome P450 (CYP), and ALDH show genetic variations (ADH1B, ALDH2, CYP2E1∗6, and CYP2E1∗7B besides others) that affect enzymatic activity in the liver and alcohol metabolism [66–68]. As a result, ethanol's pharmacokinetic and pharmacodynamic properties are affected by this genetic variation, as reflected in interracial and ethnic pharmacological differences [56, 66–71]. Consequently the risk of developing diseases may increase in certain populations, including that of hypertension [70], alcohol dependence, and several types of alcohol-related cancer [60, 72–75].
After first pass metabolism, the remaining ethanol and its metabolites are distributed in different tissue, and the excess alcohol is excreted in the breath, urine, and sweat [56]. The distribution of ethanol throughout the body is driven in direct proportion to water content of each tissue, especially at the ethanol steady-state. Since ethanol is a small, polar molecule, the distribution volume of ethanol is dependent on the total body water of an individual (50 to 60% lean body weight) [76–78]. The variation in the distribution volume of ethanol has been evaluated for women and men, and in both sexes, the distribution volume decreases as the body mass index increases [79].
Alcohol-driven physiological changes, such as vascular effects (vasodilation) or changes in cardiac output, can also modify tissue blood flow and ethanol distribution [78]. Since the blood flow to the brain remains relatively constant, changes in the blood concentration of ethanol are the most relevant factor influencing the amount of ethanol delivered to the brain and therefore for the different levels of brain intoxication [78–80].
The distribution of ethanol is also particularly relevant during pregnancy, as 1-2 hours after maternal alcohol ingestion the fetal alcohol concentrations reach levels that are nearly equivalent to the maternal levels [81]. The elimination of ethanol by the fetus is impaired due to its reduced metabolic capacity. Thus, fetal exposure is prolonged through the reuptake of amniotic-fluid containing ethanol [81]. Ultimately, the elimination of alcohol from the fetus relies on the mother's metabolic capacity, which inevitably is a process that occurs late, meaning that the fetus is exposed to the toxicological effects of alcohol [82]. Therefore, many of the physical effects of ethanol on brain structure not only affect neurobehavioral features during fetal development but may also persist into childhood, potentially enduring until adulthood [82, 83].
3.3. Ethanol and Acetate Can Reach the Brain
Ethanol can cross the blood-brain barrier and it can be metabolized in the brain. Indeed, ethanol has been found in the human brain after alcohol intake [84], although metabolites of ethanol, like acetate, can also reach the brain as products of first pass metabolism [85]. Recently, the metabolism of [2-(13)C]-ethanol was evaluated in the brains of rats, and products such as labeled acetate, glutamate, glutamine, and GABA were detected found [86].
4. Metabolism of Ethanol and Acetate in the Brain
The oxidation of ethanol to acetaldehyde can occur in the brain through pathways that involve catalase, cytochrome CYP2E1, and ADH. The main pathway to metabolize ethanol in the liver is that involving ADH, although it has not been definitively shown to play a role in ethanol metabolism in the brain. In certain regions of the adult rat, mouse, and human brain it has been possible to identify ADH mRNA transcripts, with ADH1 and ADH4 expressed at distinct sites [87, 88], yet with no detectable activity after exposure to ethanol. Nonetheless, ADH4 inhibition avoids the synaptic dysfunction associated with severe alcohol intoxication in the hippocampus [89]. Moreover ADH activity (ADH1, ADH3, and ADH4) was found in the human brain but under pathological process like brain cancer [73] and Alzheimer's disease [90], and not induced by alcohol intake. In addition, and despite fulfilling a less prominent role in ethanol metabolism [85, 91], ADHs have been related to enhanced voluntary alcohol intake in rats [92].
Other pathways metabolize ethanol in the brain. Catalase and CYP2E1 are the main pathways; there is evidence that they do indeed play an important role in ethanol oxidation to acetaldehyde in the brain [91]. Indeed, acetaldehyde production in the brain in vivo depends on catalase activity [85, 93] and catalase appears to be expressed in all neural cells. Peroxisomal catalase is a tetrameric, heme-containing enzyme that, in addition to converting hydrogen peroxide (H2O2) to water and oxygen, can also oxidize ethanol to acetaldehyde. The discovery of the catalase pathway for acetaldehyde formation in the brain represented an important first step in our understanding of the role of acetaldehyde in the effects of ethanol in the brain [94]. Studies using inhibitors of catalase and acatalasemic mice revealed that catalase is responsible for approximately half of the ethanol metabolism occurring in the CNS [91]. Indeed, inhibitors of catalase are also effective in inhibiting the production of acetaldehyde.
The cytochrome P450 enzymes (CYP2E1) that are involved in ethanol metabolism in the liver have also been implicated in its metabolism in the brain. CYP2E1 reduces molecular oxygen to water and thus ethanol is oxidized to acetaldehyde. This enzyme is induced in response to chronic drinking and it may contribute to the increased rates of ethanol elimination in heavy drinkers. Some endogenous substrates for CYP2E1 include acetone and fatty acids, both of which are abundant in the brain [95]. The CYP2E1 system fulfills an important role in the generation of reactive oxygen species (ROS) and exposure to ethanol is related to the accumulation of ROS, which in rat brain homogenates may be attributed to the induction of CYP2E1 [96]. Not only ethanol but many other substrates are also metabolized by CYP2E1, including neurotoxins or procarcinogens, producing reactive intermediates [97, 98]. Moreover, in human neurons CYP2E1 is known to generate ROS and nitric oxide through the induction of NADPH/xanthine oxidase and nitric oxide synthase [99].
Therefore, CYP2E1 and catalase are the main pathways in the brain that metabolize ethanol to acetaldehyde, while ADH appears to play a minor role. Acetaldehyde is a biologically active compound and it has been implicated in alcohol addiction [100, 101], as well as inducing euphoria at low concentrations [102]. The effects of ethanol are modulated by acetaldehyde [100, 103], which in turn may react with endogenous substances to form other biologically active compounds. Acetaldehydes along with other proteins (adducts) were found in mice brain after alcohol consumption and in alcoholic human brains, suggesting they are involved in neural damage [104, 105]. Moreover adducts like salsolinol (formed when acetaldehyde binds to dopamine) were also seen to be involved in neurotoxicity [106] and in reinforcing addictive ethanol conduct [107]. Salsolinol has been identified in the brain and cerebrospinal fluid of patients with Parkinson disease, and it has been proposed to increase ROS production along with a reduction of glutathione [108], as well as reducing intracellular ATP and thereby acting as an inhibitor of mitochondrial energy supply. Thus, acetaldehyde reinforces its own effects or enhances the addictive action of ethanol [109, 110].
As a result, acetaldehyde oxidation is required for detoxification and it can be metabolized to acetate by ALDH [111]. ALDH is critically important and the risk of alcohol-induced toxicity in individuals with mutant ALDH2 increases remarkably [112], while ALDH2 overexpression diminishes alcohol-related ROS production [113]. However, the accumulation of NADH increases in association with ALDH activity [114] and if the NAD+/NADH ratio decreases, the amounts of superoxide radicals increase [115, 116]. Moreover, although ALDH activity has beneficial effects, such as in the reduction of acetaldehyde, it also produces free radicals. Finally, the acetate produced by ALDH is metabolized in the Krebs cycle to produce energy or provide intermediaries for other molecules. Recent research showed that oxidation of [13]C-acetate generates specific neurotransmitters, as [13]C-glutamine, glutamate, and GABA levels were higher in chronic ethanol-exposed rats than in controls [86]. The production of these molecules may be related to the known effects of GABA receptors [16, 17, 19, 117], although other receptors are also involved in the effects of ethanol, such as dopamine, acetylcholine, and NMDA receptors [118–120] (Figure 2).
5. Oxidative Stress Produced by Ethanol
ROS are produced by exposure to ethanol [85] and they are associated with the effects of ethanol in the brain [92, 99, 101, 121–125], where ROS-related damage is due to oxidative stress [99, 124, 126–128]. The oxidative balance is a result of the amount of ROS that accumulates and the activity of antioxidant enzymes. In the brain, antioxidant enzymes are present in the cortex, cerebellum, hypothalamus, striatum, and spinal cord, and they include glutathione peroxidase, superoxide dismutase, glutathione reductase, and peroxiredoxin [129]. When the oxidative balance is disturbed, oxidative stress develops that affects the cell as a whole, as well as proteins, lipids, and DNA individually, provoking neurotoxicity or neurodegeneration.
6. The Antioxidant System and the Effects of Ethanol
The formation of ROS accompanies many physiological processes, such that the body has developed a system of antioxidant protection against their harmful effects. In the brain, where the generation of free radicals is particularly severe, it is essential that the antioxidant system functions correctly [130]. Antioxidant activity is considered as enzymatic or nonenzymatic based on the mechanism of action involved.
6.1. Superoxide Dismutase (SOD)
It is an enzyme that catalyzes the dismutation of the superoxide anion to hydrogen peroxide, which is then decomposed by catalases primarily located in the peroxisomes. There are two main SOD isoenzymes found in the CNS of mammals: Mn-SOD (dependent on mitochondrial manganese ions) and Cu, Zn-SOD (SOD-1) present in the cytoplasm, microsomes, and synaptosomes [131]. Increased SOD activity is considered to be an adaptive response to oxidative stress, such as that induced by acute ethanol toxicity in the cerebral cortex [132]. However, acute ethanol intoxication reduces the activity of Cu, Zn-SOD in the cytosolic and microsomal fraction of the rat brain, and Mn-SOD activity in the mitochondria [131]. SOD interacts closely with catalase, which catalyzes the deprotonation of peroxide hydrogen and the oxidation of substances like methanol, ethanol, formate, nitrite, and quinones.
6.2. Catalase
In mammals, catalase is primarily located in the liver, erythrocytes, kidneys, and CNS. In the CNS, it can be found in microsomes [133] and it has been shown that, in acute ethanol poisoning, there is an increase of catalase activity in the cytosol, microsomes, and synaptosomes, as well as a reduction in the mitochondria of the rat CNS [131]. The increase in catalase activity following ethanol intake and its effects in the CNS are associated with weak ADH activity. This increase in catalase activity in the CNS may be adaptive processes induced by the increase in the hydrogen peroxide generated, as what occurs in the CNS of animals exposed to high concentrations of ethanol [134].
6.3. Glutathione Peroxidase (GSH-Px)
It is present in many tissues, as well as in the neurons and glia of the CNS [135, 136]. The role of GSH-Px is limited to the reduction of peroxides in which glutathione participates, which is accompanied by the formation of glutathione disulfide. In the rat and human CNS, the greatest glutathione peroxidase activity is observed in the gray and white matter of the cerebral cortex [137, 138].
6.4. Glutathione Reductase (GRed)
It is an enzyme present in the cytosol and in the mitochondria of most cells, catalyzing the regeneration of reduced glutathione oxidation at the expense of NADPH. Most GRed activity is found in neurons and glial cells [139], and acute ethanol poisoning significantly dampens GRed activity in the cerebral cortex [140].
The activity of antioxidant enzymes is significantly altered in the CNS of animals chronically intoxicated with ethanol. The antioxidative capacity of the CNS also depends on exogenous antioxidants obtained by the organism through its dietary intake. The most important exogenous antioxidant in the CNS is vitamin E, and both vitamin E and vitamin C content in the CNS falls after ethanol consumption, whereas vitamin A content increases [131].
7. Oxidized Fatty Acids as a Consequence of Oxidative Stress
Lipid peroxidation affects polyunsaturated fatty acids in membrane phospholipids as oxidative stress increases, producing bioactive aldehydes like 4-hydroxyalkenals and malondialdehyde [141]. Oxidative stress and the products of lipid peroxidation, 4-hydroxynonenal (HNE) [99, 142–145] or malondialdehyde [141, 146, 147], have been related to decreased neuronal viability in some studies. Ethanol-induced lipoperoxidation by oxidative stress [142] and its products decrease the intracellular reduced glutathione and increase its oxidized form [148]. HNE has also been associated with increases in mitochondrial permeability and cytochrome c release [143, 149, 150], the latter triggering apoptotic cell death by activating caspases [145, 150]. Interestingly, the toxicity mediated by the product of lipoperoxidation was weaker when glutathione transferase A4-4 activity was enhanced and glutathionyl-HNE was produced, avoiding the accumulation of HNE [150, 151] and possibly serving as a mechanism of tolerance. However, the activation of glutathione transferase A4-4 was suppressed in the presence of anionic phospholipids like cardiolipin [152]. Furthermore, the ability of HNE to produce glutathionyl-HNE was prevented by a PLA2 inhibitor [153], suggesting a role of PLA2 in the production of HNE.
8. The Role of Phospholipids in Stress Damage
Cardiolipin is a phospholipid and it is the major component of mitochondrial membranes, although ethanol-induced oxidative stress provokes a loss of this lipid [152, 154–157] in conjunction with the appearance of HNE [157, 158]. Therefore, cardiolipin oxidation occurs following ethanol ingestion and consequently its fatty acids are released from phospholipids by PLA2. When cardiolipin is affected by ethanol, mitochondrial function is impaired and the outer mitochondrial membrane may disintegrate [157, 159], which could induce the release of cytochrome c from the mitochondria and trigger an apoptotic cascade mediated by caspases [158, 160]. Interestingly, the neurodegeneration induced by ethanol can be prevented by an inhibitor of PLA2 in vitro [153, 161].
Phosphatidylserine (PS) has also been shown to play a role in apoptotic signaling, and both the reduction in PS and the enhanced neuronal cell death that ensues during the developmental period may contribute to the brain defects often observed in fetal alcohol syndrome [162]. Meanwhile, docosahexaenoic acid (DHA: 22:6n-3) prevents neuronal apoptosis by promoting PS accumulation [162], while conversely, PLA2 activity and oxidation-mediated HNE production may diminish the levels of PS.
9. Ceramide Related to Neurodegeneration
Ceramides are produced in the central nervous system by de novo synthesis or sphingomyelin hydrolysis [163]. Ceramide has been shown to accumulate in mitochondria upon the induction of apoptotic processes related to neurodegeneration [164–175]. The expression of serine palmitoyltransferase was localized in neurons and it was enhanced in caspase 3-positive neurons induced by ethanol [172], indicating that de novo ceramide synthesis participates in ethanol-induced apoptotic neurodegeneration in the brain. Although ceramide synthase 6 (CerS6) fulfills a protective role, this enzyme produces C16-ceramides and they are the precursors of other sphingolipids, such as sphingomyelin and glucosylceramide. Interestingly, CerS6 is enhanced within hours of ethanol withdrawal as a compensatory effect [176]. In summary, ceramide is an apoptotic signal [173] but it is also necessary for the sphingomyelin synthesis required to produce diacylglycerol (DAG), which in turn activates PKC [177], thereby avoiding apoptosis [178].
10. Lipids Potentially Involved in the Compensatory Mechanisms Protecting against Ethanol-Induced Damage
While some lipids are altered to signal cells for destruction, others seem to offset some of the effects that occur due to oxidation. For example, there is more cholesterol in neuron membranes exposed to ethanol [155]. Cholesterol is known to provide rigidity to membranes and ethanol is effective in disrupting unstable lipid membranes. Hence, an increase in the cholesterol present in membranes may represent a compensatory mechanism to combat ethanol damage. Indeed, when mitochondrial cardiolipin is oxidized and its fatty acid released, membranes become unstable due to a loss of rigidity.
Other lipids can also reduce the availability or the effects of metabolites of ethanol, such as phosphatidylethanolamine, phosphatidylethanol, and acylethanolamine. Ethanol exposure augments the amount of phosphatidylethanolamine due to the attachment of aminated ethanol to citidyldiphosphate [152, 179], resulting in the production of phosphatidylethanolamine through the Kennedy pathway [180]. Moreover, phosphatidylethanolamine can serve as a substrate for acyltransferases and indeed N-acylphosphoethanolamine (NAPE) is produced following ethanol exposure [168]. The amount of NAPE in membranes augments under cellular stress and as a result of tissue damage [181–184], and NAPE represents a precursor of the N-acylethanolamines [185] involved in learning and memory [186], neuroinflammation [187], oxidative stress, neuroprotection, and neurogenesis. Palmitoylethanolamine treatment of cultured cells produces neuroprotection against oxidative stress, impeding apoptosis [187–189] and protection in mice with chronic constriction injury [190]. Moreover, the endocannabinoid anandamide is also involved in neurodegeneration and thus acylethanolamines, and especially palmitoylethanolamine, appear to play an important role as neuroprotectors. Acylethanolamines can be found in the mitochondria in vitro [191] and palmitoyl requires carnitine to enter mitochondria. When cells or animals receive carnitine it acts as a neuroprotective agent, preventing ethanol-induced damage [147]. Furthermore, ϖ type-3 unsaturated fatty acids and DHA provide neuroprotection in conjunction with an increase in the formation of acylethanolamine [161, 162], suggesting that the formation of the latter prevents the damage caused by the oxidative metabolism of ethanol. Finally, ethanol can also be metabolized as phosphatidylethanol, a molecule found in the brain of rats [192] that is possibly formed to avoid ethanol oxidation.
11. Conclusions
Lipid metabolism is clearly affected by exposure to ethanol (Figure 3), and the alterations to lipid components like cardiolipin and some phospholipids in response to ethanol provide evidence of cell damage. The formation of oxidized species, abnormal lipids, and dysfunctional membranes due to ethanol uptake also provokes cell degeneration. However, compensatory mechanisms exist to dampen the effects of these metabolic events and to minimize cell damage, as reflected by the neuroprotective activities of natural lipids like DHA, esters, vitamin E, and so forth. Thus, ethanol-induced neurodegeneration is at least partly the result of the equilibrium maintained between the toxicity of signaling lipids and the protection they confer on the cell (Figure 4).
Acknowledgment
The authors are grateful to Dr. Mark Sefton for his revisions and comments on the final version of the paper.
Conflict of Interests
The authors have no conflict of interests to declare.
References
- 1.Chartier K. G., Hesselbrock M. N., Hesselbrock V. M. Ethnicity and gender comparisons of health consequences in adults with alcohol dependence. Substance Use and Misuse. 2013;48(3):200–210. doi: 10.3109/10826084.2013.747743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Kiecolt K. J., Aggen S. H., Kendler K. S. Genetic and environmental influences on the relationship between mastery and alcohol dependence. Alcoholism: Clinical and Experimental Research. 2013;37(6):905–913. doi: 10.1111/acer.12058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Alaux-Cantin S., Warnault V., Legastelois R., et al. Alcohol intoxications during adolescence increase motivation for alcohol in adult rats and induce neuroadaptations in the nucleus accumbens. Neuropharmacology. 2013;67:521–531. doi: 10.1016/j.neuropharm.2012.12.007. [DOI] [PubMed] [Google Scholar]
- 4.Roerecke M., Rehm J. Cause-specific mortality risk in alcohol use disorder treatment patients: a systematic review and meta-analysis. International Journal of Epidemiology. 2014;43(3):906–919. doi: 10.1093/ije/dyu018.dyu018 [DOI] [PubMed] [Google Scholar]
- 5.Calhoun F., Attilia M. L., Spagnolo P. A., Rotondo C., Mancinelli R., Ceccanti M. National Institute on Alcohol Abuse and Alcoholism and the study of fetal alcohol spectrum disorders. The International Consortium. Annali dell'Istituto Superiore di Sanita. 2006;42(1):4–7. [PubMed] [Google Scholar]
- 6.Dias D., Mendonça M. C., Real F. C., Vieira D. N., Teixeira H. M. Suicides in the Centre of Portugal: seven years analysis. Forensic Science International. 2014;234(1):22–28. doi: 10.1016/j.forsciint.2013.10.034. [DOI] [PubMed] [Google Scholar]
- 7.Wickizer T. M. State-level estimates of the economic costs of alcohol and drug abuse. Journal of Health Care Finance. 2013;39(3):71–84. [PubMed] [Google Scholar]
- 8.Scally G. Crunch time for the government on alcohol pricing in England: backtracking on the minimum unit price pledge would be a public health disaster. British Medical Journal. 2013;346(7900) doi: 10.1136/bmj.f1784.f1784 [DOI] [PubMed] [Google Scholar]
- 9.Mundt M. P., Parthasarathy S., Chi F. W., Sterling S., Campbell C. I. 12-Step participation reduces medical use costs among adolescents with a history of alcohol and other drug treatment. Drug and Alcohol Dependence. 2012;126(1-2):124–130. doi: 10.1016/j.drugalcdep.2012.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Rehm J., Mathers C., Popova S., Thavorncharoensap M., Teerawattananon Y., Patra J. Global burden of disease and injury and economic cost attributable to alcohol use and alcohol-use disorders. The Lancet. 2009;373(9682):2223–2233. doi: 10.1016/S0140-6736(09)60746-7. [DOI] [PubMed] [Google Scholar]
- 11.Jedema H. P., Carter M. D., Dugan B. P., Gurnsey K., Olsen A. S., Bradberry C. W. The acute impact of ethanol on cognitive performance in rhesus macaques. Cerebral Cortex. 2011;21(8):1783–1791. doi: 10.1093/cercor/bhq244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Badawy A. A.-B., Morgan C. J., Lowett J. W. T., Bradley D. M., Thomas R. Decrease in circulating tryptophan availability to the brain after acute ethanol consumption by normal volunteers: implications for alcohol-induced aggressive behaviour and depression. Pharmacopsychiatry. 1995;28(supplement 2):93–97. doi: 10.1055/s-2007-979626. [DOI] [PubMed] [Google Scholar]
- 13.Badawy A. A.-B., Doughrty D. M., Marsh-Richard D. M., Steptoe A. Activation of liver tryptophan pyrrolase mediates the decrease in tryptophan availability to the brain after acute alcohol consumption by normal subjects. Alcohol and Alcoholism. 2009;44(3):267–271. doi: 10.1093/alcalc/agp005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Loeber S., Duka T. Extinction learning of stimulus reward contingencies: the acute effects of alcohol. Drug and Alcohol Dependence. 2009;102(1–3):56–62. doi: 10.1016/j.drugalcdep.2009.01.014. [DOI] [PubMed] [Google Scholar]
- 15.Dougherty D. M., Marsh-Richard D. M., Hatzis E. S., Nouvion S. O., Mathias C. W. A test of alcohol dose effects on multiple behavioral measures of impulsivity. Drug and Alcohol Dependence. 2008;96(1-2):111–120. doi: 10.1016/j.drugalcdep.2008.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Haughey H. M., Ray L. A., Finan P., Villanueva R., Niculescu M., Hutchison K. E. Human γ-aminobutyric acid A receptor α2 gene moderates the acute effects of alcohol and brain mRNA expression. Genes, Brain and Behavior. 2008;7(4):447–454. doi: 10.1111/j.1601-183x.2007.00369.x. [DOI] [PubMed] [Google Scholar]
- 17.Wadleigh A., Valenzuela C. F. Ethanol increases GABAergic transmission and excitability in cerebellar molecular layer interneurons from GAD67-GFP knock-in mice. Alcohol and Alcoholism. 2012;47(1):1–8. doi: 10.1093/alcalc/agr147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Tiwari V., Veeraiah P., Subramaniam V., Patel A. B. Differential effects of ethanol on regional glutamatergic and GABAergic neurotransmitter pathways in mouse brain. Journal of Neurochemistry. 2014;128(5):628–640. doi: 10.1111/jnc.12508. [DOI] [PubMed] [Google Scholar]
- 19.Wu G., Liu H., Jin J., et al. Ethanol attenuates sensory stimulus-evoked responses in cerebellar granule cells via activation of GABAA receptors in vivo in mice. Neuroscience Letters. 2014;561:107–111. doi: 10.1016/j.neulet.2013.12.049. [DOI] [PubMed] [Google Scholar]
- 20.Halliday H. L., Reid M. M., McClure G. Results of heavy drinking in pregnancy. British Journal of Obstetrics and Gynaecology. 1982;89(11):892–895. doi: 10.1111/j.1471-0528.1982.tb05052.x. [DOI] [PubMed] [Google Scholar]
- 21.West J. R., Blake C. A. Fetal alcohol syndrome: an assessment of the field. Experimental Biology and Medicine. 2005;230(6):354–356. doi: 10.1177/15353702-0323006-02. [DOI] [PubMed] [Google Scholar]
- 22.Berman R. F., Hannigan J. H. Effects of prenatal alcohol exposure on the hippocampus: spatial behavior, electrophysiology, and neuroanatomy. Hippocampus. 2000;10(1):94–110. doi: 10.1002/(sici)1098-1063(2000)10:1<94::aid-hipo11>3.0.co;2-t. [DOI] [PubMed] [Google Scholar]
- 23.Lindquist D. H., Sokoloff G., Milner E., Steinmetz J. E. Neonatal ethanol exposure results in dose-dependent impairments in the acquisition and timing of the conditioned eyeblink response and altered cerebellar interpositus nucleus and hippocampal CA1 unit activity in adult rats. Alcohol. 2013;47(6):447–457. doi: 10.1016/j.alcohol.2013.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Ramezani A., Goudarzi I., Lashkarboluki T., Ghorbanian M. T., Abrari K., Salmani M. E. Role of oxidative stress in ethanol-induced neurotoxicity in the developing cerebellum. Iranian Journal of Basic Medical Sciences. 2012;15(4):965–974. [PMC free article] [PubMed] [Google Scholar]
- 25.Thomas J. D., Abou E. J., Dominguez H. D. Prenatal choline supplementation mitigates the adverse effects of prenatal alcohol exposure on development in rats. Neurotoxicology and Teratology. 2009;31(5):303–311. doi: 10.1016/j.ntt.2009.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Krishnamoorthy M., Gerwe B. A., Scharer C. D., et al. Low ethanol concentration alters CHRNA5 RNA levels during early human development. Reproductive Toxicology. 2010;30(3):489–492. doi: 10.1016/j.reprotox.2010.04.011. [DOI] [PubMed] [Google Scholar]
- 27.Nardelli A., Lebel C., Rasmussen C., Andrew G., Beaulieu C. Extensive deep gray matter volume reductions in children and adolescents with fetal alcohol spectrum disorders. Alcoholism: Clinical and Experimental Research. 2011;35(8):1404–1417. doi: 10.1111/j.1530-0277.2011.01476.x. [DOI] [PubMed] [Google Scholar]
- 28.Kuehn D., Aros S., Cassorla F., et al. A prospective cohort study of the prevalence of growth, facial, and central nervous system abnormalities in children with heavy prenatal alcohol exposure. Alcoholism: Clinical and Experimental Research. 2012;36(10):1811–1819. doi: 10.1111/j.1530-0277.2012.01794.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Tapert S. F., Pulido C., Paulus M. P., Schuckit M. A., Burke C. Level of response to alcohol and brain response during visual working memory. Journal of Studies on Alcohol. 2004;65(6):692–700. doi: 10.15288/jsa.2004.65.692. [DOI] [PubMed] [Google Scholar]
- 30.Clark D. B., Thatcher D. L., Tapert S. F. Alcohol, psychological dysregulation, and adolescent brain development. Alcoholism: Clinical and Experimental Research. 2008;32(3):375–385. doi: 10.1111/j.1530-0277.2007.00601.x. [DOI] [PubMed] [Google Scholar]
- 31.Tapert S. F., Brown G. G., Baratta M. V., Brown S. A. fMRI BOLD response to alcohol stimuli in alcohol dependent young women. Addictive Behaviors. 2004;29(1):33–50. doi: 10.1016/j.addbeh.2003.07.003. [DOI] [PubMed] [Google Scholar]
- 32.Tapert S. F., Schweinsburg A. D., Barlett V. C., et al. Blood oxygen level dependent response and spatial working memory in adolescents with alcohol use disorders. Alcoholism: Clinical and Experimental Research. 2004;28(10):1577–1586. doi: 10.1097/01.ALC.0000141812.81234.A6. [DOI] [PubMed] [Google Scholar]
- 33.White A. M. What happened? Alcohol, memory blackouts, and the brain. Alcohol Research and Health. 2003;27(2):186–196. [PMC free article] [PubMed] [Google Scholar]
- 34.He X., Sullivan E. V., Stankovic R. K., Harper C. G., Pfefferbaum A. Interaction of thiamine deficiency and voluntary alcohol consumption disrupts rat corpus callosum ultrastructure. Neuropsychopharmacology. 2007;32(10):2207–2216. doi: 10.1038/sj.npp.1301332. [DOI] [PubMed] [Google Scholar]
- 35.Magrys S. A., Olmstead M. C. Alcohol intoxication alters cognitive skills mediated by frontal and temporal brain regions. Brain and Cognition. 2014;85:271–276. doi: 10.1016/j.bandc.2013.12.010. [DOI] [PubMed] [Google Scholar]
- 36.Zorumski C. F., Mennerick S., Izumi Y. Acute and chronic effects of ethanol on learning-related synaptic plasticity. Alcohol. 2014;48(1):1–17. doi: 10.1016/j.alcohol.2013.09.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Erritzoe D., Tziortzi A., Bargiela D., et al. In vivo imaging of cerebral dopamine D3 receptors in alcoholism. Neuropsychopharmacology. 2014;39(7):1703–1712. doi: 10.1038/npp.2014.18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Burnett E. J., Grant K. A., Davenport A. T., Hemby S. E., Friedman D. P. The effects of chronic ethanol self-administration on hippocampal 5-HT1A receptors in monkeys. Drug and Alcohol Dependence. 2014;136:135–142. doi: 10.1016/j.drugalcdep.2014.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Vetreno R. P., Hall J. M., Savage L. M. Alcohol-related amnesia and dementia: animal models have revealed the contributions of different etiological factors on neuropathology, neurochemical dysfunction and cognitive impairment. Neurobiology of Learning and Memory. 2011;96(4):596–608. doi: 10.1016/j.nlm.2011.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Vetreno R. P., Ramos R. L., Anzalone S., Savage L. M. Brain and behavioral pathology in an animal model of Wernicke's encephalopathy and Wernicke-Korsakoff syndrome. Brain Research. 2012;1436:178–192. doi: 10.1016/j.brainres.2011.11.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Akai J., Akai K. Neuropathological study of the nucleus basalis of meynert in alcoholic dementia. Arukoru Kenkyuto Yakubutsu Ison. 1989;24(2):80–88. [PubMed] [Google Scholar]
- 42.Zhang Z., Liu Y., Zhou B., et al. Altered functional connectivity of the marginal division in Alzheimer's disease. Current Alzheimer Research. 2014;11(2):145–155. doi: 10.2174/1567205011666140110112608. [DOI] [PubMed] [Google Scholar]
- 43.Andersen B. B. Reduction of Purkinje cell volume in cerebellum of alcoholics. Brain Research. 2004;1007(1-2):10–18. doi: 10.1016/j.brainres.2004.01.058. [DOI] [PubMed] [Google Scholar]
- 44.Halliday G., Cullen K., Harding A. Neuropathological correlates of memory dysfunction in the Wernicke-Korsakoff syndrome. Alcohol and Alcoholism Supplement. 1994;2:245–251. [PubMed] [Google Scholar]
- 45.Jochum T., Boettger M. K., Burkhardt C., Juckel G., Bär K.-J. Increased pain sensitivity in alcohol withdrawal syndrome. European Journal of Pain. 2010;14(7):713–718. doi: 10.1016/j.ejpain.2009.11.008. [DOI] [PubMed] [Google Scholar]
- 46.Mizuguchi M., Tomonaga M., Fukusato T., Asano M. Acute necrotizing encephalopathy with widespread edematous lesions of symmetrical distribution. Acta Neuropathologica. 1989;78(1):108–111. doi: 10.1007/BF00687411. [DOI] [PubMed] [Google Scholar]
- 47.Riethdorf L., Warzok R., Schwesinger G. Alcoholic encephalopathies in autopsy material. Zentralblatt für Pathologie. 1991;137(1):48–56. [PubMed] [Google Scholar]
- 48.Ramchandani V. A., Bosron W. F., Li T. K. Research advances in ethanol metabolism. Pathologie Biologie. 2001;49(9):676–682. doi: 10.1016/s0369-8114(01)00232-2. [DOI] [PubMed] [Google Scholar]
- 49.Levitt M. D., Li R., Demaster E. G., Elson M., Furne J., Levitt D. G. Use of measurements of ethanol absorption from stomach and intestine to assess human ethanol metabolism. American Journal of Physiology—Gastrointestinal and Liver Physiology. 1997;273(4):G951–G957. doi: 10.1152/ajpgi.1997.273.4.G951. [DOI] [PubMed] [Google Scholar]
- 50.Levitt M. D., Fine C., Furne J. K., Levitt D. G. Use of maltose hydrolysis measurements to characterize the interaction between the aqueous diffusion barrier and the epithelium in the rat jejunum. The Journal of Clinical Investigation. 1996;97(10):2308–2315. doi: 10.1172/jci118673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Pang K. S. Modeling of intestinal drug absorption: roles of transporters and metabolic enzymes (for the Gillette review series) Drug Metabolism and Disposition. 2003;31(12):1507–1519. doi: 10.1124/dmd.31.12.1507. [DOI] [PubMed] [Google Scholar]
- 52.Oneta C. M., Simanowski U. A., Martinez M., et al. First pass metabolism of ethanol is strikingly influenced by the speed of gastric emptying. Gut. 1998;43(5):612–619. doi: 10.1136/gut.43.5.612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Shultz J., Weiner H., Westcott J. Retardation of ethanol absorption by food in the stomach. Journal of Studies on Alcohol. 1980;41(9):861–870. doi: 10.15288/jsa.1980.41.861. [DOI] [PubMed] [Google Scholar]
- 54.Sedman A. J., Wilkinson P. K., Sakmar E., Weidler D. J., Wagner J. G. Food effects on absorption and metabolism of alcohol. Journal of Studies on Alcohol. 1976;37(9):1197–1214. doi: 10.15288/jsa.1976.37.1197. [DOI] [PubMed] [Google Scholar]
- 55.Lee S.-L., Chau G.-Y., Yao C.-T., Wu C.-W., Yin S.-J. Functional assessment of human alcohol dehydrogenase family in ethanol metabolism: significance of first-pass metabolism. Alcoholism: Clinical and Experimental Research. 2006;30(7):1132–1142. doi: 10.1111/j.1530-0277.2006.00139.x. [DOI] [PubMed] [Google Scholar]
- 56.Norberg Å., Jones W. A., Hahn R. G., Gabrielsson J. L. Role of variability in explaining ethanol pharmacokinetics: research and forensic applications. Clinical Pharmacokinetics. 2003;42(1):1–31. doi: 10.2165/00003088-200342010-00001. [DOI] [PubMed] [Google Scholar]
- 57.Deetz J. S., Luehr C. A., Vallee B. L. Human liver alcohol dehydrogenase isozymes: reduction of aldehydes and ketones. Biochemistry. 1984;23(26):6822–6828. doi: 10.1021/bi00321a084. [DOI] [PubMed] [Google Scholar]
- 58.Haseba T., Tomita Y., Kurosu M., Ohno Y. Dose and time changes in liver alcohol dehydrogenase (ADH) activity during acute alcohol intoxication involve not only class I but also class III ADH and govern elimination rate of blood ethanol. Legal Medicine. 2003;5(4):202–211. doi: 10.1016/S1344-6223(03)00080-4. [DOI] [PubMed] [Google Scholar]
- 59.Haseba T., Kameyama K., Mashimo K., Ohno Y. Dose-dependent change in elimination kinetics of ethanol due to shift of dominant metabolizing enzyme from ADH 1 (Class I) to ADH 3 (Class III) in mouse. International Journal of Hepatology. 2012;2012:8. doi: 10.1155/2012/408190.408190 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Li T.-K., Yin S.-J., Crabb D. W., O'Connor S., Ramchandani V. A. Genetic and environmental influences on alcohol metabolism in humans. Alcoholism: Clinical and Experimental Research. 2001;25(1):136–144. doi: 10.1111/j.1530-0277.2001.tb02138.x. [DOI] [PubMed] [Google Scholar]
- 61.Smolen T. N., Smolen A., van de Kamp J. L. Developmental profile of hepatic alcohol and aldehyde dehydrogenase activities in long-sleep and short-sleep mice. Alcohol. 1990;7(1):69–74. doi: 10.1016/0741-8329(90)90062-H. [DOI] [PubMed] [Google Scholar]
- 62.Lundquist F., Tygstrup N., Winkler K., Mellemgaard K., Munck-Petersen S. Ethanol metabolism and production of free acetate in the human liver. The Journal of Clinical Investigation. 1962;41:955–961. doi: 10.1172/jci104574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Lindros K. O., Stowell A., Pikkarainen P., Salaspuro M. Elevated blood acetaldehyde in alcoholics with accelerated ethanol elimination. Pharmacology Biochemistry and Behavior. 1980;13(supplement 1):119–124. doi: 10.1016/s0091-3057(80)80019-0. [DOI] [PubMed] [Google Scholar]
- 64.Puig J. G., Fox I. H. Ethanol-induced activation of adenine nucleotide turnover. Evidence for a role of acetate. The Journal of Clinical Investigation. 1984;74(3):936–941. doi: 10.1172/jci111512. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Sarkola T., Iles M. R., Kohlenberg-Mueller K., Eriksson C. J. P. Ethanol, acetaldehyde, acetate, and lactate levels after alcohol intake in white men and women: effect of 4-methylpyrazole. Alcoholism: Clinical and Experimental Research. 2002;26(2):239–245. doi: 10.1111/j.1530-0277.2002.tb02530.x. [DOI] [PubMed] [Google Scholar]
- 66.Kayaalti Z., Soylemezoglu T. Distribution of ADH1B, ALDH2, CYP2E1 ∗6, and CYP2E1 ∗7B genotypes in Turkish population. Alcohol. 2010;44(5):415–423. doi: 10.1016/j.alcohol.2010.06.002. [DOI] [PubMed] [Google Scholar]
- 67.Montano Loza A. J., Ramirez Iglesias M. T., Perez Diaz I., et al. Association of alcohol-metabolizing genes with alcoholism in a Mexican Indian (Otomi) population. Alcohol. 2006;39(2):73–79. doi: 10.1016/j.alcohol.2006.07.001. [DOI] [PubMed] [Google Scholar]
- 68.García-Bañuelos J., Panduro A., Gordillo-Bastidas D., et al. Genetic polymorphisms of genes coding to alcohol-metabolizing enzymes in Western Mexicans: association of CYP2E1∗c2/CYP2E1∗5B allele with cirrhosis and liver function. Alcoholism: Clinical and Experimental Research. 2012;36(3):425–431. doi: 10.1111/j.1530-0277.2011.01617.x. [DOI] [PubMed] [Google Scholar]
- 69.Seng K. Y., Limenta L. M. G., Heng D., Lee E. J. D. Population pharmacokinetics and pharmacogenetics of alcohol in Chinese and Indians in Singapore. Journal of Clinical Pharmacy and Therapeutics. 2013;38(2):141–149. doi: 10.1111/jcpt.12003. [DOI] [PubMed] [Google Scholar]
- 70.Yamada Y., Sun F., Tsuritani I., Honda R. Genetic differences in ethanol metabolizing enzymes and blood pressure in Japanese alcohol consumers. Journal of Human Hypertension. 2002;16(7):479–486. doi: 10.1038/sj.jhh.1001415. [DOI] [PubMed] [Google Scholar]
- 71.Lorenzo A., Auguet T., Vidal F., et al. Polymorphisms of alcohol-metabolizing enzymes and the risk for alcoholism and alcoholic liver disease in Caucasian Spanish women. Drug and Alcohol Dependence. 2006;84(2):195–200. doi: 10.1016/j.drugalcdep.2006.03.002. [DOI] [PubMed] [Google Scholar]
- 72.Chen Y.-C., Peng G.-S., Wang M.-F., Tsao T.-P., Yin S.-J. Polymorphism of ethanol-metabolism genes and alcoholism: correlation of allelic variations with the pharmacokinetic and pharmacodynamic consequences. Chemico-Biological Interactions. 2009;178(1–3):2–7. doi: 10.1016/j.cbi.2008.10.029. [DOI] [PubMed] [Google Scholar]
- 73.Laniewska-Dunaj M., Jelski W., Orywal K., Kochanowicz J., Rutkowski R., Szmitkowski M. The activity of class I, II, III and IV of alcohol dehydrogenase (ADH) isoenzymes and aldehyde dehydrogenase (ALDH) in brain cancer. Neurochemical Research. 2013;38(7):1517–1521. doi: 10.1007/s11064-013-1053-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Sanchez-Alvarez R., Martinez-Outschoorn U. E., Lin Z., et al. Ethanol exposure induces the cancer-associated fibroblast phenotype and lethal tumor metabolism: implications for breast cancer prevention. Cell Cycle. 2013;12(2):289–301. doi: 10.4161/cc.23109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Chan I. S., Guy C. D., Machado M. V., et al. Alcohol activates the hedgehog pathway and induces related procarcinogenic processes in the alcohol-preferring rat model of hepatocarcinogenesis. Alcoholism: Clinical and Experimental Research. 2014;38(3):787–800. doi: 10.1111/acer.12279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Cowan J. M., Jr., Weathermon A., McCutcheon J. R., Oliver R. D. Determination of volume of distribution for ethanol in male and female subjects. Journal of Analytical Toxicology. 1996;20(5):287–290. doi: 10.1093/jat/20.5.287. [DOI] [PubMed] [Google Scholar]
- 77.Endres H. G. E., Grüner O. Comparison of D2O and ethanol dilutions in total body water measurements in humans. The Clinical Investigator. 1994;72(11):830–837. doi: 10.1007/bf00190736. [DOI] [PubMed] [Google Scholar]
- 78.Pizon A. F., Becker C. E., Bikin D. The clinical significance of variations in ethanol toxicokinetics. Journal of Medical Toxicology. 2007;3(2):63–72. doi: 10.1007/bf03160911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Maudens K. E., Patteet L., van Nuijs A. L. N., Van Broekhoven C., Covaci A., Neels H. The influence of the Body Mass Index (BMI) on the volume of distribution of ethanol. Forensic Science International. 2014;243:74–78. doi: 10.1016/j.forsciint.2014.04.036. [DOI] [PubMed] [Google Scholar]
- 80.Marxen M., Gan G., Schwarz D., et al. Acute effects of alcohol on brain perfusion monitored with arterial spin labeling magnetic resonance imaging in young adults. Journal of Cerebral Blood Flow and Metabolism. 2014;34(3):472–479. doi: 10.1038/jcbfm.2013.223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Burd L., Blair J., Dropps K. Prenatal alcohol exposure, blood alcohol concentrations and alcohol elimination rates for the mother, fetus and newborn. Journal of Perinatology. 2012;32(9):652–659. doi: 10.1038/jp.2012.57. [DOI] [PubMed] [Google Scholar]
- 82.Moore E. M., Riley E. P. What happens when children with fetal alcohol spectrum disorders become adults? Current Developmental Disorders Reports. 2015;2(3):219–227. doi: 10.1007/s40474-015-0053-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Skorput A. G., Gupta V. P., Yeh P. W., Yeh H. H. Persistent interneuronopathy in the prefrontal cortex of young adult offspring exposed to ethanol in utero . The Journal of Neuroscience. 2015;35(31):10977–10988. doi: 10.1523/jneurosci.1462-15.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Fein G., Meyerhoff D. J. Ethanol in human brain by magnetic resonance spectroscopy: correlation with blood and breath levels, relaxation, and magnetization transfer. Alcoholism: Clinical and Experimental Research. 2000;24(8):1227–1235. doi: 10.1111/j.1530-0277.2000.tb02088.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Zimatkin S. M., Buben A. I. Ethanol oxidation in the living brain. Alcohol and Alcoholism. 2007;42(6):529–532. doi: 10.1093/alcalc/agm059. [DOI] [PubMed] [Google Scholar]
- 86.Wang J., Du H., Ma X., et al. Metabolic products of [2–13C]ethanol in the rat brain after chronic ethanol exposure. Journal of Neurochemistry. 2013;127(3):353–364. doi: 10.1111/jnc.12405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Martínez S. E., Vaglenova J., Sabrià J., Carmen Martínez M., Farrés J., Parés X. Distribution of alcohol dehydrogenase mRNA in the rat central nervous system. Consequences for brain ethanol and retinoid metabolism. European Journal of Biochemistry. 2001;268(19):5045–5056. doi: 10.1046/j.0014-2956.2001.02416.x. [DOI] [PubMed] [Google Scholar]
- 88.Galter D., Carmine A., Buervenich S., Duester G., Olson L. Distribution of class I, III and IV alcohol dehydrogenase mRNAS in the adult rat, mouse and human brain. European Journal of Biochemistry. 2003;270(6):1316–1326. doi: 10.1046/j.1432-1033.2003.03502.x. [DOI] [PubMed] [Google Scholar]
- 89.Tokuda K., Izumi Y., Zorumski C. F. Locally-generated acetaldehyde contributes to the effects of ethanol on neurosteroids and LTP in the hippocampus. Neurology and Clinical Neuroscience. 2013;1(4):138–147. doi: 10.1111/ncn3.39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Borger E., Aitken L., Du H., Zhang W., Gunn-Moore F. J., Du Yan S. S. Is amyloid binding alcohol dehydrogenase a drug target for treating Alzheimer's disease? Current Alzheimer Research. 2013;10(1):21–29. [PMC free article] [PubMed] [Google Scholar]
- 91.Zimatkin S. M., Pronko S. P., Vasiliou V., Gonzalez F. J., Deitrich R. A. Enzymatic mechanisms of ethanol oxidation in the brain. Alcoholism: Clinical and Experimental Research. 2006;30(9):1500–1505. doi: 10.1111/j.1530-0277.2006.00181.x. [DOI] [PubMed] [Google Scholar]
- 92.Karahanian E., Quintanilla M. E., Tampier L., et al. Ethanol as a prodrug: brain metabolism of ethanol mediates its reinforcing effects. Alcoholism: Clinical and Experimental Research. 2011;35(4):606–612. doi: 10.1111/j.1530-0277.2011.01439.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Rhoads D. E., Contreras C., Fathalla S. Brain levels of catalase remain constant through strain, developmental, and chronic alcohol challenges. Enzyme Research. 2012;2012:6. doi: 10.1155/2012/572939.572939 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Aragon C. M. G., Rogan F., Amit Z. Ethanol metabolism in rat brain homogenates by a catalase-H2O2 system. Biochemical Pharmacology. 1992;44(1):93–98. doi: 10.1016/0006-2952(92)90042-h. [DOI] [PubMed] [Google Scholar]
- 95.Lieber C. S. Microsomal ethanol-oxidizing system (MEOS): the first 30 years (1968–1998)—a review. Alcoholism: Clinical and Experimental Research. 1999;23(6):991–1007. [PubMed] [Google Scholar]
- 96.Zhong Y., Dong G., Luo H., et al. Induction of brain CYP2E1 by chronic ethanol treatment and related oxidative stress in hippocampus, cerebellum, and brainstem. Toxicology. 2012;302(2-3):275–284. doi: 10.1016/j.tox.2012.08.009. [DOI] [PubMed] [Google Scholar]
- 97.Koop D. R. Oxidative and reductive metabolism by cytochrome P450 2E1. The FASEB Journal. 1992;6(2):724–730. doi: 10.1096/fasebj.6.2.1537462. [DOI] [PubMed] [Google Scholar]
- 98.Leung T., Rajendran R., Singh S., Garva R., Krstic-Demonacos M., Demonacos C. Cytochrome P450 2E1 (CYP2E1) regulates the response to oxidative stress and migration of breast cancer cells. Breast Cancer Research. 2013;15(6, article R107) doi: 10.1186/bcr3574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Haorah J., Ramirez S. H., Floreani N., Gorantla S., Morsey B., Persidsky Y. Mechanism of alcohol-induced oxidative stress and neuronal injury. Free Radical Biology and Medicine. 2008;45(11):1542–1550. doi: 10.1016/j.freeradbiomed.2008.08.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Deehan G. A., Jr., Hauser S. R., Wilden J. A., Truitt W. A., Rodd Z. A. Elucidating the biological basis for the reinforcing actions of alcohol in the mesolimbic dopamine system: the role of active metabolites of alcohol. Frontiers in Behavioral Neuroscience. 2013;7, article 104 doi: 10.3389/fnbeh.2013.00104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Deng X.-S., Deitrich R. A. Putative role of brain acetaldehyde in ethanol addiction. Current Drug Abuse Reviews. 2008;1(1):3–8. doi: 10.2174/1874473710801010003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Eriksson C. J. The role of acetaldehyde in the actions of alcohol (update 2000) Alcoholism Clinical and Experimental Research. 2001;25(5, supplement):15s–32s. doi: 10.1097/00000374-200105051-00005. [DOI] [PubMed] [Google Scholar]
- 103.Cederbaum A. I. Alcohol metabolism. Clinics in Liver Disease. 2012;16(4):667–685. doi: 10.1016/j.cld.2012.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Nakamura K., Iwahashi K., Furukawa A., et al. Acetaldehyde adducts in the brain of alcoholics. Archives of Toxicology. 2003;77(10):591–593. doi: 10.1007/s00204-003-0465-8. [DOI] [PubMed] [Google Scholar]
- 105.Nakamura K., Iwahashi K., Itoh M., et al. Immunohistochemical study on acetaldehyde adducts in alcohol-fed mice. Alcoholism, Clinical and Experimental Research. 2000;24(4, supplement):93S–96S. [PubMed] [Google Scholar]
- 106.Xie G., Krnjević K., Ye J.-H. Salsolinol modulation of dopamine neurons. Frontiers in Behavioral Neuroscience. 2013;7, article 52 doi: 10.3389/fnbeh.2013.00052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Quintanilla M. E., Rivera-Meza M., Berrios-Cárcamo P. A., et al. Salsolinol, free of isosalsolinol, exerts ethanol-like motivational/sensitization effects leading to increases in ethanol intake. Alcohol. 2014;48(6):551–559. doi: 10.1016/j.alcohol.2014.07.003. [DOI] [PubMed] [Google Scholar]
- 108.Wanpen S., Govitrapong P., Shavali S., Sangchot P., Ebadi M. Salsolinol, a dopamine-derived tetrahydroisoquinoline, induces cell death by causing oxidative stress in dopaminergic SH-SY5Y cells, and the said effect is attenuated by metallothionein. Brain Research. 2004;1005(1-2):67–76. doi: 10.1016/j.brainres.2004.01.054. [DOI] [PubMed] [Google Scholar]
- 109.Peana A. T., Acquasb E. Behavioral and biochemical evidence of the role of acetaldehyde in the motivational effects of ethanol. Frontiers in Behavioral Neuroscience. 2013;7, article 86 doi: 10.3389/fnbeh.2013.00086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Quintanilla M. E., Rivera-Meza M., Berríos-Cárcamo P., Cassels B. K., Herrera-Marschitz M., Israel Y. (R)-Salsolinol, a product of ethanol metabolism, stereospecifically induces behavioral sensitization and leads to excessive alcohol intake. Addiction Biology. 2015 doi: 10.1111/adb.12268. [DOI] [PubMed] [Google Scholar]
- 111.Kikonyogo A., Pietruszko R. Aldehyde dehydrogenase from adult human brain that dehydrogenates γ-aminobutyraldehyde: purification, characterization, cloning and distribution. Biochemical Journal. 1996;316(part 1):317–324. doi: 10.1042/bj3160317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Zhang Y., Ren J. ALDH2 in alcoholic heart diseases: molecular mechanism and clinical implications. Pharmacology and Therapeutics. 2011;132(1):86–95. doi: 10.1016/j.pharmthera.2011.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Bai J., Mei Y. Overexpression of aldehyde dehydrogenase-2 attenuates neurotoxicity induced by 4-hydroxynonenal in cultured primary hippocampal neurons. Neurotoxicity Research. 2011;19(3):412–422. doi: 10.1007/s12640-010-9183-1. [DOI] [PubMed] [Google Scholar]
- 114.Manzo-Avalos S., Saavedra-Molina A. Cellular and mitochondrial effects of alcohol consumption. International Journal of Environmental Research and Public Health. 2010;7(12):4281–4304. doi: 10.3390/ijerph7124281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Antosiewicz J., Spodnik J. H., Teranishi M., et al. NADH-generating substrates reduce peroxyl radical toxicity in RL-34 cells. Folia Morphologica. 2009;68(4):247–255. [PubMed] [Google Scholar]
- 116.Velayutham M., Hemann C., Zweier J. L. Removal of H2O2 and generation of superoxide radical: role of cytochrome c and NADH. Free Radical Biology and Medicine. 2011;51(1):160–170. doi: 10.1016/j.freeradbiomed.2011.04.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Walker B. M., Koob G. F. The γ-aminobutyric acid-B receptor agonist baclofen attenuates responding for ethanol in ethanol-dependent rats. Alcoholism: Clinical and Experimental Research. 2007;31(1):11–18. doi: 10.1111/j.1530-0277.2006.00259.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Rex E. B., Rankin M. L., Ariano M. A., Sibley D. R. Ethanol regulation of D1 dopamine receptor signaling is mediated by protein kinase C in an isozyme-specific manner. Neuropsychopharmacology. 2008;33(12):2900–2911. doi: 10.1038/npp.2008.16. [DOI] [PubMed] [Google Scholar]
- 119.Hendrickson L. M., Gardner P., Tapper A. R. Nicotinic acetylcholine receptors containing the α4 subunit are critical for the nicotine-induced reduction of acute voluntary ethanol consumption. Channels. 2011;5(2):124–127. doi: 10.4161/chan.5.2.14409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Villegas E., Estruch R., Mengod G., Cortés R. NMDA receptors in frontal cortex and hippocampus of alcohol consumers. Addiction Biology. 2011;16(1):163–165. doi: 10.1111/j.1369-1600.2009.00201.x. [DOI] [PubMed] [Google Scholar]
- 121.Correa M., Arizzi M. N., Betz A., Mingote S., Salamone J. D. Open field locomotor effects in rats after intraventricular injections of ethanol and the ethanol metabolites acetaldehyde and acetate. Brain Research Bulletin. 2003;62(3):197–202. doi: 10.1016/j.brainresbull.2003.09.013. [DOI] [PubMed] [Google Scholar]
- 122.Volkow N. D., Kim S. W., Wang G.-J., et al. Acute alcohol intoxication decreases glucose metabolism but increases acetate uptake in the human brain. NeuroImage. 2013;64(1):277–283. doi: 10.1016/j.neuroimage.2012.08.057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Corsini G. U., Zuddas A., Bonuccelli U., Schinelli S., Kopin I. J. 1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) neurotoxicity in mice is enhanced by ethanol or acetaldehyde. Life Sciences. 1987;40(9):827–832. doi: 10.1016/0024-3205(87)90030-0. [DOI] [PubMed] [Google Scholar]
- 124.Heap L., Ward R. J., Abiaka C., et al. The influence of brain acetaldehyde on oxidative status, dopamine metabolism and visual discrimination task. Biochemical Pharmacology. 1995;50(2):263–270. doi: 10.1016/0006-2952(94)00539-X. [DOI] [PubMed] [Google Scholar]
- 125.Wan J. Y., Wang J. Y., Wang Y., Wang J. Y. A comparison between acute exposures to ethanol and acetaldehyde on neurotoxicity, nitric oxide production and NMDA-induced excitotoxicity in primary cultures of cortical neurons. The Chinese Journal of Physiology. 2000;43(3):131–138. [PubMed] [Google Scholar]
- 126.Crews F. T., Nixon K. Mechanisms of neurodegeneration and regeneration in alcoholism. Alcohol and Alcoholism. 2009;44(2):115–127. doi: 10.1093/alcalc/agn079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Ikonomidou C., Kaindl A. M. Neuronal death and oxidative stress in the developing brain. Antioxidants and Redox Signaling. 2011;14(8):1535–1550. doi: 10.1089/ars.2010.3581. [DOI] [PubMed] [Google Scholar]
- 128.Luo J. Mechanisms of ethanol-induced death of cerebellar granule cells. Cerebellum. 2012;11(1):145–154. doi: 10.1007/s12311-010-0219-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Shim S.-Y., Kim H.-S. Oxidative stress and the antioxidant enzyme system in the developing brain. Korean Journal of Pediatrics. 2013;56(3):107–111. doi: 10.3345/kjp.2013.56.3.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Augustyniak A., Michalak K., Skrzydlewska E. The action of oxidative stress induced by ethanol on the central nervous system (CNS) Postępy Higieny i Medycyny Doświadczalnej. 2005;59:464–471. [PubMed] [Google Scholar]
- 131.Reddy S. K., Husain K., Schlorff E. C., Scott R. B., Somani S. M. Dose response of ethanol ingestion on antioxidant defense system in rat brain subcellular fractions. NeuroToxicology. 1999;20(6):977–987. [PubMed] [Google Scholar]
- 132.Enache M., Van Waes V., Vinner E., Lhermitte M., Maccari S., Darnaudéry M. Impact of an acute exposure to ethanol on the oxidative stress status in the hippocampus of prenatal restraint stress adolescent male rats. Brain Research. 2008;1191:55–62. doi: 10.1016/j.brainres.2007.11.031. [DOI] [PubMed] [Google Scholar]
- 133.Hipolito L., Sanchez M. J., Polache A., Granero L. Brain metabolism of ethanol and alcoholism: an update. Current Drug Metabolism. 2007;8(7):716–727. doi: 10.2174/138920007782109797. [DOI] [PubMed] [Google Scholar]
- 134.Fukui Y., Sakata-Haga H. Intrauterine environment-genome interaction and children's development (1): ethanol: a teratogen in developing brain. Journal of Toxicological Sciences. 2009;34(supplement 2):SP273–SP278. doi: 10.2131/jts.34.sp273. [DOI] [PubMed] [Google Scholar]
- 135.Bosch-Morell F., Martínez-Soriano F., Colell A., Fernández-Checa J. C., Romero F. J. Chronic ethanol feeding induces cellular antioxidants decrease and oxidative stress in rat peripheral nerves. Effect of S-adenosyl-l-methionine and N-acetyl-l-cysteine. Free Radical Biology and Medicine. 1998;25(3):365–368. doi: 10.1016/S0891-5849(98)00036-7. [DOI] [PubMed] [Google Scholar]
- 136.Chalimoniuk M., Jagsz S., Sadowska-Krepa E., Chrapusta S. J., Klapcinska B., Langfort J. Diversity of endurance training effects on antioxidant defenses and oxidative damage in different brain regions of adolescent male rats. Journal of Physiology and Pharmacology. 2015;66(4):539–547. [PubMed] [Google Scholar]
- 137.Ansari K. A., Bigelow D., Kaplan E. Glutathione peroxidase activity in surgical and autopsied human brains. Neurochemical Research. 1985;10(5):703–711. doi: 10.1007/bf00964408. [DOI] [PubMed] [Google Scholar]
- 138.Huang J., Philbert M. A. Distribution of glutathione and glutathione-related enzyme systems in mitochondria and cytosol of cultured cerebellar astrocytes and granule cells. Brain Research. 1995;680(1-2):16–22. doi: 10.1016/0006-8993(95)00209-9. [DOI] [PubMed] [Google Scholar]
- 139.Dringen R., Hirrlinger J. Glutathione pathways in the brain. Biological Chemistry. 2003;384(4):505–516. doi: 10.1515/bc.2003.059. [DOI] [PubMed] [Google Scholar]
- 116.Somani S. M., Husain K. Interaction of exercise training and chronic ethanol ingestion on antioxidant system of rat brain regions. Journal of Applied Toxicology. 1997;17(5):329–336. doi: 10.1002/(sici)1099-1263(199709)17:5<329::aid-jat452>3.0.co;2-t. [DOI] [PubMed] [Google Scholar]
- 141.Pizzimenti S., Ciamporcero E., Daga M., et al. Interaction of aldehydes derived from lipid peroxidation and membrane proteins. Frontiers in Physiology. 2013;4, article 242 doi: 10.3389/fphys.2013.00242. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Chen J. J., Schenker S., Henderson G. I. 4-Hydroxynonenal levels are enhanced in fetal liver mitochondria by in utero ethanol exposure. Hepatology. 1997;25(1):142–147. doi: 10.1002/hep.510250126. [DOI] [PubMed] [Google Scholar]
- 143.Ramachandran V., Perez A., Chen J., Senthil D., Schenker S., Henderson G. I. In utero ethanol exposure causes mitochondrial dysfunction, which can result in apoptotic cell death in fetal brain: a potential role for 4-hydroxynonenal. Alcoholism: Clinical and Experimental Research. 2001;25(6):862–871. doi: 10.1111/j.1530-0277.2001.tb02292.x. [DOI] [PubMed] [Google Scholar]
- 144.Ramachandran V., Watts L. T., Maffi S. K., Chen J., Schenker S., Henderson G. I. Ethanol-induced oxidative stress precedes mitochondrially mediated apoptotic death of cultured fetal cortical neurons. Journal of Neuroscience Research. 2003;74(4):577–588. doi: 10.1002/jnr.10767. [DOI] [PubMed] [Google Scholar]
- 145.Maffi S. K., Rathinam M. L., Cherian P. P., et al. Glutathione content as a potential mediator of the vulnerability of cultured fetal cortical neurons to ethanol-induced apoptosis. Journal of Neuroscience Research. 2008;86(5):1064–1076. doi: 10.1002/jnr.21562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Rendón-Ramírez A., Cortés-Couto M., Martínez-Rizo A. B., Muñiz-Hernández S., Velázquez-Fernández J. B. Oxidative damage in young alcohol drinkers: a preliminary study. Alcohol. 2013;47(7):501–504. doi: 10.1016/j.alcohol.2013.08.002. [DOI] [PubMed] [Google Scholar]
- 147.Augustyniak A., Skrzydlewska E. The influence of L-carnitine suplementation on the antioxidative abilities of serum and the central nervous system of ethanol-induced rats. Metabolic Brain Disease. 2010;25(4):381–389. doi: 10.1007/s11011-010-9217-7. [DOI] [PubMed] [Google Scholar]
- 148.Liu W., Kato M., Akhand A. A., et al. 4-hydroxynonenal induces a cellular redox status-related activation of the caspase cascade for apoptotic cell death. Journal of Cell Science. 2000;113(part 4):635–641. doi: 10.1242/jcs.113.4.635. [DOI] [PubMed] [Google Scholar]
- 149.Mahr A., Batteux F., Tubiana S., et al. Brief report: prevalence of antineutrophil cytoplasmic antibodies in infective endocarditis. Arthritis & Rheumatology. 2014;66(6):1672–1677. doi: 10.1002/art.38389. [DOI] [PubMed] [Google Scholar]
- 150.Raza H., John A. 4-Hydroxynonenal induces mitochondrial oxidative stress, apoptosis and expression of glutathione S-transferase A4-4 and cytochrome P450 2E1 in PC12 cells. Toxicology and Applied Pharmacology. 2006;216(2):309–318. doi: 10.1016/j.taap.2006.06.001. [DOI] [PubMed] [Google Scholar]
- 151.McElhanon K. E., Bose C., Sharma R., Wu L., Awasthi Y. C., Singh S. P. 4 null mouse embryonic fibroblasts exhibit enhanced sensitivity to oxidants: role of 4-hydroxynonenal in oxidant toxicity. Open Journal of Apoptosis. 2013;2(1):1–11. doi: 10.4236/ojapo.2013.21001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Shimoji M., Imaizumi N., Aniya Y. Modulation of membrane-bound glutathione transferase activity by phospholipids including cardiolipin. Biological and Pharmaceutical Bulletin. 2011;34(2):209–213. doi: 10.1248/bpb.34.209. [DOI] [PubMed] [Google Scholar]
- 153.Moon K.-H., Tajuddin N., Brown J., Neafsey E. J., Kim H.-Y., Collins M. A. Phospholipase A2, oxidative stress, and neurodegeneration in binge ethanol-treated organotypic slice cultures of developing rat brain. Alcoholism: Clinical and Experimental Research. 2014;38(1):161–169. doi: 10.1111/acer.12221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Rodrigues M. A. C., Rodrigues J. L., Martins N. M., et al. Carvedilol protects against cisplatin-induced oxidative stress, redox state unbalance and apoptosis in rat kidney mitochondria. Chemico-Biological Interactions. 2011;189(1-2):45–51. doi: 10.1016/j.cbi.2010.10.014. [DOI] [PubMed] [Google Scholar]
- 155.Barceló-Coblijn G., Wold L. E., Ren J., Murphy E. J. Prenatal ethanol exposure increases brain cholesterol content in adult rats. Lipids. 2013;48(11):1059–1068. doi: 10.1007/s11745-013-3821-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Reddy V. D., Padmavathi P., Kavitha G., Saradamma B., Varadacharyulu N. Alcohol-induced oxidative/nitrosative stress alters brain mitochondrial membrane properties. Molecular and Cellular Biochemistry. 2013;375(1-2):39–47. doi: 10.1007/s11010-012-1526-1. [DOI] [PubMed] [Google Scholar]
- 157.Petrosillo G., Portincasa P., Grattagliano I., et al. Mitochondrial dysfunction in rat with nonalcoholic fatty liver. Involvement of complex I, reactive oxygen species and cardiolipin. Biochimica et Biophysica Acta (BBA)—Bioenergetics. 2007;1767(10):1260–1267. doi: 10.1016/j.bbabio.2007.07.011. [DOI] [PubMed] [Google Scholar]
- 158.Liu W., Porter N. A., Schneider C., Brash A. R., Yin H. Formation of 4-hydroxynonenal from cardiolipin oxidation: intramolecular peroxyl radical addition and decomposition. Free Radical Biology and Medicine. 2011;50(1):166–178. doi: 10.1016/j.freeradbiomed.2010.10.709. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Rubin E., Rottenberg H. Ethanol-induced injury and adaptation in biological membranes. Federation Proceedings. 1982;41(8):2465–2471. [PubMed] [Google Scholar]
- 160.Vladimirov Y. A., Proskurnina E. V., Alekseev A. V. Molecular mechanisms of apoptosis. Structure of cytochrome c-cardiolipin complex. Biochemistry. 2013;78(10):1086–1097. doi: 10.1134/s0006297913100027. [DOI] [PubMed] [Google Scholar]
- 161.Brown J., III, Achille N., Neafsey E. J., Collins M. A. Binge ethanol-induced neurodegeneration in rat organotypic brain slice cultures: effects of PLA2 inhibitor mepacrine and Docosahexaenoic Acid (DHA) Neurochemical Research. 2009;34(2):260–267. doi: 10.1007/s11064-008-9765-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Akbar M., Baick J., Calderon F., Wen Z., Kim H.-Y. Ethanol promotes neuronal apoptosis by inhibiting phosphatidylserine accumulation. Journal of Neuroscience Research. 2006;83(3):432–440. doi: 10.1002/jnr.20744. [DOI] [PubMed] [Google Scholar]
- 163.Horres C. R., Hannun Y. A. The roles of neutral sphingomyelinases in neurological pathologies. Neurochemical Research. 2012;37(6):1137–1149. doi: 10.1007/s11064-011-0692-y. [DOI] [PubMed] [Google Scholar]
- 164.Pascual M., Valles S. L., Renau-Piqueras J., Guerri C. Ceramide pathways modulate ethanol-induced cell death in astrocytes. Journal of Neurochemistry. 2003;87(6):1535–1545. doi: 10.1046/j.1471-4159.2003.02130.x. [DOI] [PubMed] [Google Scholar]
- 165.Saito M., Saito M., Cooper T. B., Vadasz C. Ethanol-induced changes in the content of triglycerides, ceramides, and glucosylceramides in cultured neurons. Alcoholism: Clinical and Experimental Research. 2005;29(8):1374–1383. doi: 10.1097/01.alc.0000175011.22307.61. [DOI] [PubMed] [Google Scholar]
- 166.Schatter B., Jin S., Löffelholz K., Klein J. Cross-talk between phosphatidic acid and ceramide during ethanol-induced apoptosis in astrocytes. BMC Pharmacology. 2005;5, article 3 doi: 10.1186/1471-2210-5-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Dasgupta S., Adams J. A., Hogan E. L. Maternal alcohol consumption increases sphingosine levels in the brains of progeny mice. Neurochemical Research. 2007;32(12):2217–2224. doi: 10.1007/s11064-007-9445-3. [DOI] [PubMed] [Google Scholar]
- 168.Saito M., Chakraborty G., Mao R.-F., et al. Ethanol alters lipid profiles and phosphorylation status of AMP-activated protein kinase in the neonatal mouse brain. Journal of Neurochemistry. 2007;103(3):1208–1218. doi: 10.1111/j.1471-4159.2007.04836.x. [DOI] [PubMed] [Google Scholar]
- 169.Babenko N. A., Shakhova E. G. Effects of flavonoids on sphingolipid turnover in the toxin-damaged liver and liver cells. Lipids in Health and Disease. 2008;7, article 1 doi: 10.1186/1476-511x-7-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Novgorodov S. A., Gudz T. I., Obeid L. M. Long-chain ceramide is a potent inhibitor of the mitochondrial permeability transition pore. The Journal of Biological Chemistry. 2008;283(36):24707–24717. doi: 10.1074/jbc.m801810200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.de La Monte S. M., Longato L., Tong M., Denucci S., Wands J. R. The liver-brain axis of alcohol-mediated neurodegeneration: role of toxic lipids. International Journal of Environmental Research and Public Health. 2009;6(7):2055–2075. doi: 10.3390/ijerph6072055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Saito M., Chakraborty G., Hegde M., et al. Involvement of ceramide in ethanol-induced apoptotic neurodegeneration in the neonatal mouse brain. Journal of Neurochemistry. 2010;115(1):168–177. doi: 10.1111/j.1471-4159.2010.06913.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Wang G., Bieberich E. Prenatal alcohol exposure triggers ceramide-induced apoptosis in neural crest-derived tissues concurrent with defective cranial development. Cell Death and Disease. 2010;1(5, article e46) doi: 10.1038/cddis.2010.22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Deng T.-X., Wang Z.-X., Gao X.-Q., et al. Alcohol-induced proliferation of neurons in mouse hippocampal dentate gyrus: a possible role of ceramide. Sheng Li Xue Bao. 2011;63(6):479–490. [PubMed] [Google Scholar]
- 175.Ramirez T., Longato L., Tong M., Wands J. R., de la Monte S. M., Dostalek M. Insulin resistance, ceramide accumulation and endoplasmic reticulum stress in experimental chronic alcohol-induced steatohepatitis. Alcohol and Alcoholism. 2013;48(1):39–52. doi: 10.1093/alcalc/ags106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Bae M., Bandaru V. V. R., Patel N., Haughey N. J. Ceramide metabolism analysis in a model of binge drinking reveals both neuroprotective and toxic effects of ethanol. Journal of Neurochemistry. 2014;131(5):645–654. doi: 10.1111/jnc.12834. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Cerbón J., López-Sánchez R. D. C. Diacylglycerol generated during sphingomyelin synthesis is involved in protein kinase C activation and cell proliferation in Madin-Darby canine kidney cells. Biochemical Journal. 2003;373(3):917–924. doi: 10.1042/bj20021732. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Novgorodov S. A., Chudakova D. A., Wheeler B. W., et al. Developmentally regulated ceramide synthase 6 increases mitochondrial Ca2+ loading capacity and promotes apoptosis. The Journal of Biological Chemistry. 2011;286(6):4644–4658. doi: 10.1074/jbc.m110.164392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Wen Z., Kim H.-Y. Alterations in hippocampal phospholipid profile by prenatal exposure to ethanol. Journal of Neurochemistry. 2004;89(6):1368–1377. doi: 10.1111/j.1471-4159.2004.02433.x. [DOI] [PubMed] [Google Scholar]
- 180.Gibellini F., Smith T. K. The Kennedy pathway—de novo synthesis of phosphatidylethanolamine and phosphatidylcholine. IUBMB Life. 2010;62(6):414–428. doi: 10.1002/iub.337. [DOI] [PubMed] [Google Scholar]
- 181.Subbanna S., Shivakumar M., Psychoyos D., Xie S., Basavarajappa B. S. Anandamide-CB1 receptor signaling contributes to postnatal ethanol-induced neonatal neurodegeneration, adult synaptic, and memory deficits. The Journal of Neuroscience. 2013;33(15):6350–6366. doi: 10.1523/jneurosci.3786-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Taléns-Visconti R., Sanchez-Vera I., Kostic J., et al. Neural differentiation from human embryonic stem cells as a tool to study early brain development and the neuroteratogenic effects of ethanol. Stem Cells and Development. 2011;20(2):327–339. doi: 10.1089/scd.2010.0037. [DOI] [PubMed] [Google Scholar]
- 183.Suárez J., Romero-Zerbo Y., Márquez L., et al. Ulcerative colitis impairs the acylethanolamide-based anti-inflammatory system reversal by 5-aminosalicylic acid and glucocorticoids. PLoS ONE. 2012;7(5) doi: 10.1371/journal.pone.0037729.e37729 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Gebeh A. K., Willets J. M., Bari M., et al. Elevated anandamide and related N-acylethanolamine levels occur in the peripheral blood of women with ectopic pregnancy and are mirrored by changes in peripheral fatty acid amide hydrolase activity. Journal of Clinical Endocrinology and Metabolism. 2013;98(3):1226–1234. doi: 10.1210/jc.2012-3390. [DOI] [PubMed] [Google Scholar]
- 185.Tsuboi K., Okamoto Y., Ikematsu N., et al. Enzymatic formation of N-acylethanolamines from N-acylethanolamine plasmalogen through N-acylphosphatidylethanolamine-hydrolyzing phospholipase D-dependent and -independent pathways. Biochimica et Biophysica Acta—Molecular and Cell Biology of Lipids. 2011;1811(10):565–577. doi: 10.1016/j.bbalip.2011.07.009. [DOI] [PubMed] [Google Scholar]
- 186.Maroof N., Pardon M. C., Kendall D. A. Endocannabinoid signalling in Alzheimer's disease. Biochemical Society Transactions. 2013;41(6):1583–1587. doi: 10.1042/BST20130140. [DOI] [PubMed] [Google Scholar]
- 187.Garg P., Duncan R. S., Kaja S., Koulen P. Intracellular mechanisms of N-acylethanolamine-mediated neuroprotection in a rat model of stroke. Neuroscience. 2010;166(1):252–262. doi: 10.1016/j.neuroscience.2009.11.069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Lombardi G., Miglio G., Varsaldi F., Minassi A., Appendino G. Oxyhomologation of the amide bond potentiates neuroprotective effects of the endolipid N-palmitoylethanolamine. Journal of Pharmacology and Experimental Therapeutics. 2007;320(2):599–606. doi: 10.1124/jpet.106.112987. [DOI] [PubMed] [Google Scholar]
- 189.Duncan R. S., Chapman K. D., Koulen P. The neuroprotective properties of palmitoylethanolamine against oxidative stress in a neuronal cell line. Molecular Neurodegeneration. 2009;4(1, article 50) doi: 10.1186/1750-1326-4-50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Di Cesare Mannelli L., D'Agostino G., Pacini A., et al. Palmitoylethanolamide is a disease-modifying agent in peripheral neuropathy: pain relief and neuroprotection share a PPAR-alpha-mediated mechanism. Mediators of Inflammation. 2013;2013:12. doi: 10.1155/2013/328797.328797 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Duncan R. S., Xin H., Goad D. L., Chapman K. D., Koulen P. Protection of neurons in the retinal ganglion cell layer against excitotoxicity by the N-acylethanolamine, N-linoleoylethanolamine. Clinical Ophthalmology. 2011;5(1):543–548. doi: 10.2147/opth.s18309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Alling C., Gustavsson L., Månsson J.-E., Benthin G., Änggård E. Phosphatidylethanol formation in rat organs after ethanol treatment. Biochimica et Biophysica Acta (BBA)—Lipids and Lipid Metabolism. 1984;793(1):119–122. doi: 10.1016/0005-2760(84)90060-2. [DOI] [PubMed] [Google Scholar]