Next Article in Journal
Deciphering Interactions between Phosphorus Status and Toxic Metal Exposure in Plants and Rhizospheres to Improve Crops Reared on Acid Soil
Next Article in Special Issue
Dysregulated Signaling at Postsynaptic Density: A Systematic Review and Translational Appraisal for the Pathophysiology, Clinics, and Antipsychotics’ Treatment of Schizophrenia
Previous Article in Journal
Future Perspectives of Prime Editing for the Treatment of Inherited Retinal Diseases
Previous Article in Special Issue
Long-Term Memory Formation in Drosophila Depends on the 3′UTR of CPEB Gene orb2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synaptic Plasticity Abnormalities in Fetal Alcohol Spectrum Disorders

1
Center for Dementia Research, Nathan Kline Institute for Psychiatric Research, Orangeburg, NY 10962, USA
2
Molecular Imaging and Neuropathology Area, New York State Psychiatric Institute, New York, NY 10032, USA
3
Department of Psychiatry, Columbia University Irving Medical Center, New York, NY 10032, USA
4
Department of Psychiatry, New York University Langone Medical Center, New York, NY 10016, USA
*
Author to whom correspondence should be addressed.
Cells 2023, 12(3), 442; https://doi.org/10.3390/cells12030442
Submission received: 15 December 2022 / Revised: 10 January 2023 / Accepted: 27 January 2023 / Published: 29 January 2023
(This article belongs to the Special Issue New Advances in Synaptic Dysfunctions and Plasticity)

Abstract

:
The brain’s ability to strengthen or weaken synaptic connections is often termed synaptic plasticity. It has been shown to function in brain remodeling following different types of brain damage (e.g., drugs of abuse, alcohol use disorders, neurodegenerative diseases, and inflammatory conditions). Although synaptic plasticity mechanisms have been extensively studied, how neural plasticity can influence neurobehavioral abnormalities in alcohol use disorders (AUDs) is far from being completely understood. Alcohol use during pregnancy and its harmful effects on the developing offspring are major public health, social, and economic challenges. The significant attribute of prenatal alcohol exposure on offspring is damage to the central nervous system (CNS), causing a range of synaptic structural, functional, and behavioral impairments, collectively called fetal alcohol spectrum disorder (FASD). Although the synaptic mechanisms in FASD are limited, emerging evidence suggests that FASD pathogenesis involves altering a set of molecules involved in neurotransmission, myelination, and neuroinflammation. These studies identify several immediate and long-lasting changes using many molecular approaches that are essential for synaptic plasticity and cognitive function. Therefore, they can offer potential synaptic targets for the many neurobehavioral abnormalities observed in FASD. In this review, we discuss the substantial research progress in different aspects of synaptic and molecular changes that can shed light on the mechanism of synaptic dysfunction in FASD. Increasing our understanding of the synaptic changes in FASD will significantly advance our knowledge and could provide a basis for finding novel therapeutic targets and innovative treatment strategies.

1. Introduction

Alcohol is the most regularly abused substance in the world. A recent report from the National Survey on Drug Use and Health indicated that approximately 70% of people aged 18 or older reported alcohol use within the past year [1]. Additionally, nearly 25% of this population engaged in binge drinking (consuming five drinks or more in men and more than four drinks in women within approximately 2 h) within the last month [1]. Despite public education endeavors and recommendations to avoid alcohol use while pregnant, alcohol use by pregnant women remains widespread [2,3,4,5,6,7]. The use of alcohol during pregnancy in the USA has also soared over the years [3,7].
The impact of alcohol abuse during pregnancy on the developing fetus has received extensive attention since the first finding indicating the devastating and persistent complications of fetal alcohol exposure [8]. Among the potential negative consequences of fetal alcohol exposure, brain maturation impairments, followed by lifelong physical, emotional and behavioral problems, are prominent. The major outcome of the effects of high-dose alcohol exposure during early development is fetal alcohol syndrome (FAS) [9]. Pre- and postnatal abnormal growth, craniofacial defects and long-lasting behavioral difficulties characterize FAS. FAS is observed in 1–2 of every 1000 newborns [10]. In addition, FAS has been identified as a significant cause of nongenetic intellectual disabilities and behavioral difficulties worldwide [11,12,13,14,15]. Nonetheless, FAS is not the only enduring abnormality stemming from developmental alcohol exposure. The term fetal alcohol spectrum disorders (FASDs) [16] has been used to exemplify the range of persistent structural and behavioral impairments of developmental alcohol effects, with FAS on the extreme side of the spectrum. The prevalence of FASDs is estimated to be as high as 2–5% [17,18]. In addition, the current literature suggests that severe neuropsychological impairments, such as verbal learning/recall abilities, learning, and memory, characterize FASD [13,19,20,21,22,23]. These neuropsychological deficits have also been associated with most intellectual disabilities in the Western world [21,23,24,25] and impose many daily challenges for children with FASD. The severity of the neurobehavioral outcome of FASD differs and depends on the abuse patterns (continuous versus binge drinking), amount [26,27], and developmental timing [28]. Effective treatments for FASD are currently lacking because the mechanisms underlying alcohol-induced brain damage and neurobehavioral impairments are poorly defined. However, changes in the expression of genes and proteins [29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48] related to neuronal survival, neuronal growth and development, neurotransmission, myelination, synapse formation, and dendritic spines have been suggested to play a significant role in alcohol neurotoxicity. Notably, synapses maintain their molecular composition, plasticity, and function through protein homeostasis, and alterations in such homeostasis may lead to persistent neuronal plasticity deficits.

Molecular Basis of Neuronal Plasticity

Neuronal plasticity is an essential process that regulates neuronal activity by enabling neurons to fine-tune their synaptic strength in response to changes in activity. This ability is essential during development when circuits are fine-tuned by selective pruning and synapse remodeling in response to experience [49]. By altering synaptic strength, the nervous system can remodel itself, creating long-lasting memories that form the biological basis for brain function. Uncovering the mechanisms regulating synaptic plasticity will help to illuminate how neuronal plasticity disruptions influence the disorder’s pathophysiology, identify new therapeutic targets, and reveal potential impacts of pharmacologically targeting neuronal plasticity.
A central component of plasticity includes the temporal coexistence of activity. Supposedly, spike-timing dependent plasticity is a Hebbian learning rule in which the changes in synaptic strength rely on the relative timing of action potentials [50]. In monosynaptic pairs of neurons, if an output spike from postsynaptic neurons follows immediately after an input spike from the presynaptic neuron, that input becomes stronger, causing long-term potentiation (LTP). If the input spike arises immediately after an output spike, conversely, that input is formed weaker, causing long-term depression (LTD) [51,52,53]. Although STDP primarily involves molecular changes at the synapse, it depends on various distinct mechanisms that can vary in different brain regions, among different neurons within the same brain region, or among similar types of neurons. This critical window of timing dependency covers tens of milliseconds. It has profound consequences on the brain function, producing an activity-dependent bidirectional modification of synaptic strength and eventually establishing the physiological basis for learning and memory. The mechanism underlying STDP involves two different glutamate receptors that are frequently coexpressed, the α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPAR) and the N-methyl-D-aspartate receptor (NMDAR) [54]. The preexisting membrane depolarization of the NMDAR is sequentially mediated through the coactivation of the AMPAR [55] to establish mutual cellular mechanisms that enable long-term synaptic changes.

2. Influence of Developmental Alcohol on Neuronal Plasticity

There is growing evidence that neuronal plasticity is persistently impaired in animal models of FASD. Several laboratories have reported neuronal plasticity involving glutamatergic, GABAergic, and their modulator defects in different brain regions using various animal models of FASD and analyses. Here, we discuss the current understanding of these events, emphasizing pre- and postnatal alcohol studies that have contributed to the FASD field under each section.

2.1. Developmental Alcohol Influence on the Glutamatergic Neurotransmitter System

Glutamate facilitates most excitatory neurotransmission in the mammalian brain by binding to metabotropic glutamate receptors (mGluRs). mGluRs are G protein-coupled receptors and ionotropic glutamate receptors (iGluRs). iGluRs are cation-permeable ligand-gated ion channels. mGluR and iGluR activation results in distinct cellular responses on remarkably different time scales. iGluRs are classified into different functional classes, that is, AMPA receptors, kainate receptors, NMDARs, and GluD receptors (delta or δ receptors) [56,57] (Figure 1). In certain conditions, iGluR subtypes that classically mediate current responses (i.e., ionotropic signaling) also have the ability to promote intracellular signaling (i.e., metabotropic signaling) through different mechanisms [58,59].

2.1.1. Influence on AMPA Function

Given the multifactorial functions of iGluRs in normal brain function, it is unsurprising that their dysregulation is implicated in several pathophysiological conditions [60,61,62,63]. Indeed, in heterologous expression systems and some neuronal preparations, alcohol has been shown to inhibit AMPAR function when receptor function was studied using agonist application to isolated cells or oocytes [64,65]. Inhibition of AMPAR function was evident in acutely isolated hippocampal cells and cultured CNS neurons from postnatal Day 10–20 mice [66]. These findings suggest that alcohol (10–500 mM) inhibition of AMPAR function was due to the stabilization of receptor desensitization [66]. In another study [67], alcohol (50 mM) significantly depressed AMPA-evoked currents in hippocampal slices derived from neonates (postnatal day (PD) 3–6 but not in those from juveniles (PD21–PD26). Additionally, alcohol (50 mM) significantly reduced AMPA-mediated excitatory postsynaptic currents (EPSCs) in neurons from neonatal but not from juvenile rats. Additionally, alcohol (50 mM) significantly increased the paired-pulse plasticity of the AMPA-mediated current ratio in slices from PD5 rats. Furthermore, acute alcohol (50 mM) also decreased miniature excitatory postsynaptic current (mEPSC) frequency in neurons from neonatal rats. These studies suggest that alcohol reduced the probability of glutamate release at CA3 pyramidal neurons in neonatal rats by inhibiting N-type voltage-gated Ca2+ channels (VGCCs) [67]. In another study, alcohol (40 and 80 mM) inhibited AMPAR-mediated field excitatory postsynaptic potentials (fEPSPs) in acute coronal brain slices prepared from PD7–9 rats. Similarly, alcohol (80 mM) also reduced AMPAR-mediated fEPSPs in the presence of an inhibitor of Ca2+-permeable AMPARs. In the same study, alcohol (80 mM) inhibited the LTP of AMPAR-mediated fEPSPs [68].
Because hippocampal AMPAR-mediated synaptic transmission occurs during the first two postnatal weeks in rodents and plays a role in network formation and synapse development [69,70], the findings discussed above highlight the critical role of AMPARs in the action of alcohol during early postnatal development. In another study, acute alcohol (30 and 60 mM) depressed mEPSC frequencies in cultured hippocampal neurons derived from PD1 mice [71], indicating that alcohol inhibits glutamate release. These alcohol doses also enhanced anandamide (N-arachidonoylethanolamide, AEA) and 2-arachidonylglycerol (2-AG), two well-known endocannabinoids (eCBs) shown to depress excitatory neurons by binding to cannabinoid receptor type 1 (CB1), in cultured neurons [71]. Furthermore, the CB1 receptor antagonist (SR141716A) reversed acute alcohol-induced depression of mEPSC frequency. Additionally, drugs that enhance the in vivo function of eCBs blocked alcohol effects on mEPSC frequency. These findings indicate that acute alcohol-enhanced eCBs are responsible for alcohol-reduced glutamate release in these cultured cells from PD1 mice. Interestingly, it has been well established that CB1 receptor activation is associated with inhibition of N-type and P/Q-type Ca2+ channels [72] consisting of alcohol inhibition of N-type voltage-gated Ca2+ channels (VGCCs) [67], which may occur via CB1 receptor activation. There is also evidence indicating that alcohol inhibits glutamate release in other hippocampal regions. It has been shown that alcohol inhibits KCl-induced vesicular FMI-43 [N-(-3-trethylammoniumpropyl)-4-(-(dibutylamino) styryl) pyridinium dibromide] destaining in the CA1 stratum radiatum of PD21–28 rats, and that the effect was inhibited by blockers of N-type and P/Q-type Ca2+ channels [73]. Collectively, these findings corroborate the notion that alcohol inhibits glutamate release in pyramidal neurons by activating an eCB/CB1 receptor system that is negatively coupled to N-type and P/Q-type Ca2+ channels [71]. In line with these observations, recent studies have implicated CB1 function in developmental alcohol-induced defects. For example, CB1 receptor antagonist (SR141716A, SR) administration before postnatal alcohol exposure on postnatal Day 7 (PD7) prevented LTP defects in the hippocampus (HP) and cognitive impairments in adult mice [43,44,45,74,75]. Similarly, PD7 alcohol exposure failed to cause LTP and cognitive impairments in global CB1 null adult mice [43,44,45,74,75]. Both SR preadministration and use of global CB1 null mice prevented PD7 alcohol-induced neurodegeneration in neonatal mice [43,44,45,74,75]. These findings suggest that PD7 alcohol-activated CB1 limited NMDA receptor function, causing neurodegeneration and LTP/cognitive defects by inhibiting glutamate release [45].

2.1.2. Influence on AMPA Subunits Changes

Different developmental alcohol studies also examined AMPAR subunit changes and found different outcomes depending on the model, offspring age, and the region measured (Table 1). For example, alcohol exposure of neonatal rats 3 h a day from PD4–9 significantly reduced cortical GluR1 levels in western blot analysis [76]. Alcohol exposure from GD2–67 in pregnant Dunkin–Hartley strain guinea pigs enhanced GluR2/3 density (binding studies) in the cerebral cortex [77]. In another study, alcohol exposure throughout pregnancy in rats caused cognitive deficits in offspring which was correlated with reduced AMPA-mediated mEPSCs in the HP [78]. Furthermore, alcohol exposure from GD8–20 significantly enhanced Ca2+-permeable AMPA receptor (CP-AMPAR, GluR3) expression and enhanced the depression of AMPAR-EPSCs, caused synaptic strength and facilitated anti-Hebbian LTP in VTA-DA neurons of alcohol-exposed PD14–84 animals [79]. In addition, enhanced GluR3-positive particles in cytosol/TH-positive dendrites were observed in electron microscopy analysis [79]. Alcohol exposure restricted to between GD10–18 and PD4–14 increased AMPA receptor function in adult (PD74) medial prefrontal cortex (mPFC) layer VI pyramidal neurons [80]. In the same study, the mice exhibited attention deficits. Using the same alcohol administration regimen, alcohol enhanced AMPA receptor subunits (GluR1, GluR2, and GluR3) in the dentate gyrus (DG) region of PD77 rats [81]. In another study, voluntary alcohol exposure from GD1–PD21 in mice decreased the GluR1/GluR2 ratio in adolescent mouse offspring [82].
Furthermore, Ca2+-permeable receptors have been shown to help preserve the availability of Ca2+ ions necessary for LTP maintenance [83]. These findings suggest that an increase in Ca2+-impermeable receptors at the synapse (GluR2-containing) inhibits the probability of action potential generation and therefore affects NMDA-dependent LTP generation, unlike Ca2+-permeable (GluR2-lacking) receptors. Although these findings collectively suggest that alcohol exposure during different stages of pregnancy significantly affects AMPA receptor subunits and their function, how these changes contribute to synaptic plasticity and cognitive deficits in offspring warrants further study.

2.1.3. Influence on NMDA Receptors Subunits

NMDARs play a fundamental role in the neuronal plasticity and neurotransmission that regulate the learning and memory process, which is essential to CNS development and function and is also involved in neurotoxicity. NMDARs are Ca2+-permeable tetrameric protein complexes (Figure 1) composed of two obligatory GluN1 subunits and two GluN2 (A–D) or GluN3 (A, B) subunits [63]. Each NMDA receptor activation requires membrane depolarization as well as the binding of both glycine (to GluN1) and glutamate (to GluN2) for the opening of the channel pores [63]. In addition, the GluN2 subunits play a critical role in regulating many of the electrophysiological properties of the receptor. While it is generally agreed that alcohol acutely inhibits NMDA receptors, several studies have indicated that sensitivity to alcohol depends on developmental age and brain region [84,85].
Studies in oocytes have suggested that NMDARs containing GluN2A or GluN2B subunits are more sensitive to alcohol than those containing GluN2C/D [87,88]. Alcohol acts as a partial NMDAR antagonist that inhibits NMDAR-mediated EPSCs over a range of concentrations (1–100 mM). Nevertheless, the inhibition remains partial even at higher concentrations (100–500 mM) [67,89,90,91]. Alcohol has been shown to interact with the NMDA receptors GluN1 and GluN2 subunits, regulating both receptor kinetics and alcohol sensitivity, suggesting that the opposite location of these residues influences the ability of alcohol to control receptor activity [92,93,94]. Although these early studies indicated that alcohol directly affects NMDA receptors, the in vivo animal studies that examined NMDAR changes after alcohol exposure during early development have reported contrasting results. These outcomes depended on the animal species used (e.g., rat, mice, or guinea pig), the timing and period of ethanol exposure (during only one human equivalent trimester of gestation, throughout gestation, during pre- and postnatal periods) and ethanol exposure paradigms (gavage, liquid diet, etc.) and the age at which the outcome was evaluated (neonate, adolescent, or adult offspring). Nevertheless, a large body of literature suggests alterations in NMDAR subunit expression and changes in receptor function (Table 1). For example, a liquid diet containing 3.35% alcohol throughout the gestation period reduced NMDA-sensitive [3H]-glutamate binding site density in various subregions of the HP at 45 days of offspring age [95]. Similarly, exposure to alcohol from GD2 to GD62 (4 g/kg/day) in pregnant guinea pigs reduced [3H]MK-801 binding in the HP of fetuses (GD63) [96]. Prenatal and postnatal alcohol exposure caused a significant reduction in the density of NMDARs. It increased the percentage of high-affinity state (open channel state) relative to low-affinity state (closed channel state) receptors in the cortex and HP [97]. Alcohol exposure throughout pregnancy reduced GluN1, GluN2A, and GluN2B in the barrel field cortex of adult offspring [98]. Additionally, alcohol exposure throughout pregnancy and lactation significantly increased the HP levels of GluN1 and GluN2D mRNAs on PD7 and 14, reduced GluN2C on PD1 and increased [3H]MK-801 binding [99]. If exposure was limited to the sole postnatal period (PD4–9) in rats, GluN2A was increased in the cortex at PD21 in rats [100].
Furthermore, alcohol exposure throughout pregnancy reduced the C2-terminal variant postsynaptic (PSD95-immunoprecipitated) GluN2A subunit in PD21 rat offspring [101]. After prenatal ethanol exposure up to PD9, GluN2A was increased at PD10 in the HP without changes in the cortex [102]. In a mouse study, alcohol exposure to GD8 reduced GluN2B mRNA and increased GluN2A mRNA in PD90 offspring [103]. In another study, alcohol exposure from GD2 until GD63 increased GluN1 mRNA levels in both the CA3 and CA1 areas of the HP of near-term guinea pig fetuses (GD63) [104]. Alcohol exposure throughout pregnancy and lactation reduced GluN1 transcripts in PD60 and PD90 offspring HP [105]. Changes in NMDA receptor subunits in response to alcohol exposure during pre- and/or postnatal development in different models are summarized in Table 1.
Table 1. Summary of developmental alcohol exposure effects on glutamatergic neurotransmitter system.
Table 1. Summary of developmental alcohol exposure effects on glutamatergic neurotransmitter system.
Alcohol ModelBACTissue/RegionEffects
AMPAR function
PD4–9; (vapor, 3h/day) (W rats)330 mg/dLNCReduced GluR1 (PD10) [76].
GD2–67; Oral (4 g/kg) (Pigs)327 mg/dL, maternalCerebral CEReduced GluR3 (PD61) [77].
GD3–21; (4 g/kg/day) (SD Rats)184 mg/dL, maternalHPReduced AMPA-mediated mEPSCs [78].
GD8–20; (3 & 4 g/kg/day) (SD Rats)281–341 mg/dL, maternalVTAEnhanced Depression of AMPAR-EPSCs [78] and increased GluR3 (PD35–60).
GD1–21 (Alc liquid diet) (LE Rats)60 mg/dLHP (DG)Decreased GluR1 (PD90) [106]
GD8–20 (3 &4 g/kg/day) (SD Rats)281–341 mg/dL, maternalVTAEnhanced Ca2+-permeable AMPAR& increased GluR3 (PD14–84) [79].
GD10–11 (2 g/kg/day); GD12–18 (4 g/kg/day);
PD4–5 (1.5 g/kg/day);
PD6–14 (3 g/kg/day) (C57 mice)
255 mg/dL on PD10mPFC
DG
Increased AMPA receptor function (PD74) [80].
Increased GluR1, 2 & 3 (PD74) [81].
GD1–PD21 (10,15 & 20% Alc sol) (C57 mice)80 mg/dL, maternalHPReduced GluR1/GluR2 ratio (PD30–58) [82]
NMDAR function
GD1–21 (3.35% Alc liquid diet) (SD Rats)35–40 mg/dL, maternalDorsal HPReduced NMDA sensitive [3H]-glutamate binding (PD45) [95].
GD2–62 (4 g/kg/day) (SD Rats)269 mg/dL, maternalHPReduced [3H]MK-801 binding (GD63) [96].
GD12–18 (5 g/kg/day)
(SD Rats)
143 mg/dL, maternalHP & CEReduced [3H]MK-801 binding (PD20–22); Increased high-affinity state [97].
PD4–14 (10.2 % Alc diet) (SD rats)429 mg/dL, PD8HP & CEReduced [3H]MK-801 binding (PD20–22); Increased high-affinity state [97].
GD1–21 (6.5% Alc liquid diet (LE Rats)133 mg/dL, maternalBarrel field CEReduced GluN1, GluN2A & B (PD90) [98].
GD1–PD9 (10% Alc sol (SD Rats)86–112 mg/dL, maternalHPIncreased GluN1 & GluN2D mRNA (PD7 & 14) [99].
Reduced GluN2C (PD1) [99].
Increased [3H]MK-801 binding [99].
GD3–21 (20–36% Alc liquid diet (SD Rats)119–138 mg/dL, maternalHP and FBReduced GluN2A and GluN2B (PD14) in FB; Reduced GluN2B (PD7) in HP [107].
PD4–9 (6.2 g/kg/day)307 mg/dL (PD4–9)CEIncreased GluN2A (PD21) [100].
GD1–21 (20–36% Alc liquid diet (SD Rats)120–145 mg/dL, maternalCerebral CEReduced cell surface c2-terminal variant postsynaptic GluN2A (PD21) [101].
GD (5 g/kg) & PD4–9 (6.2 g/kg) (SD Rats)95 mg/dL, maternal; 35 mg/dL, pupHPIncreased GluN2A (PD10) [102].
GD2–67; Oral (4 g/kg) (Pigs)327 mg/dL, maternalWBReduced GluN2B (PD61) [77].
GD8; 25% alc, i.p. (C57 mice)NDWBReduced GluN2B mRNA; increased GluN2A mRNA (P90) [103].
GD2–63; Oral (4 g/kg) (Pigs)283 mg/dL, maternalHPIncreased GluN1 mRNA (GD63) [104].
GD1–PD14; 10% Alc sol (SD Rats)86–112 mg/dL, maternalHPReduced GluN1 mRNA (PD60&90) [105].
ND, not determined.
Concerning GluN1 subunit expression, some studies report it to be increased in the DG of the HP [31], whereas others did not find any changes [108]. The latter authors also reported no changes in GluN2A and GluN2B subunit expression in the DG. Others found that GluN2B was decreased at PD7 in rats or adult mice, accompanied by a decrease in PSD95–GluN2B complex association [109], whereas GluN2B increased in some regions of the prefrontal cortex (PFC) [110]. Interestingly, GluN2B expression was reduced in the DG [30]. At the same time, it was increased in CA1 [111] and caused enhanced LTD, revealing the substructure alterations induced by alcohol. Finally, learning deficits in HP-dependent tasks in adult rats after postnatal alcohol exposure were accompanied by dysregulation in HP gene expression through a significant reduction in glutamate-related genes, including those for GluN2B and GluN2D subunits [112]. Developmental alcohol influence on changes in NMDAR subunits may contribute to synapse dysfunction because the inflow of Ca2+ through this receptor is highly dependent upon its subunit composition [61]. In particular, the amount of Ca2+ inflow across NMDARs is higher in GluN2B-containing receptors than in GluN2A-containing receptors. Additionally, bidirectional synaptic plasticity is dependent on the postsynaptic Ca2+ concentration [113]. Alcohol exposure using a liquid diet from GD3 until birth caused no change in the forebrain PSD-95-associated NMDA-receptor complex in PD1 rats [114]. Reduced whole brain GluN2B mRNA was reported in adult offspring exposed to alcohol on GD8 [103]. In another study, alcohol exposure (two bottled choices) throughout pregnancy decreased PSD-95-associated GluN2B levels in adult HP. Although the mechanism is unclear, these findings suggest that synaptic GluN2B-containing NMDA receptor concentrations decreased in gestational alcohol-exposed adult offspring [109]. Reduced GluN2B-containing NMDA receptors can affect extracellular signal-regulated kinase (ERK) phosphorylation [115], which was reduced in the HP of alcohol-exposed adult offspring [116]. As activation of the Erk1/2 signaling pathway through NMDA receptors [117] is required for various forms of neuronal plasticity [118,119], these observations have relevance in developmental alcohol-induced LTP deficits [120,121] and HP-dependent learning and memory impairments [122].

2.1.4. Influence on NMDA Receptor Functions

Disruption of the PSD-95/GluN2B complex has been shown to reduce cAMP response element-binding protein CREB (a known target of ERK) phosphorylation [123]. Alcohol exposure from GD1 to GD21 caused reduced DG LTP in adolescent males but increased DG LTP in adolescent females [124]. A similar alcohol exposure paradigm also reduced DG LTP in adult males but not in female offspring [108]. These findings indicate that gestational alcohol might have induced neurophysiological alterations during cortical development with increased number and function of NMDARs in females and decreased in males. Consistent with this notion, GD1 to GD21 alcohol exposure enhanced glutamine synthetase expression in the DG with modified excitatory neurotransmission in exposed offspring [108]. In contrast, mice exposed to alcohol throughout pregnancy exhibited a larger NMDA-eEPSC amplitude in the orbital frontal cortex region of female adult offspring. At the same time, reduced NMDA-eEPSC current density was observed in male offspring. In the same study, the contribution of GluN2B subunit-containing NMDARs to eEPSCs was not altered by alcohol. In addition, no change in GluN2B expression in the synaptic fraction of alcohol-exposed males and females was reported [125]. Alcohol exposure throughout the gestational period reduced GluN2B subunit levels and impaired NMDAR-dependent LTP in the DG [30]. Similarly, alcohol exposure from GD3 until PD7 impaired NMDAR-dependent LTP in the CA1 and NMDAR-mediated LTP in the DG of offspring (PD21–60). Moreover, alcohol-exposed offspring rats displayed increased NMDAR-mediated transmission in both HP areas [126]. PD7 alcohol exposure in CD1 mice reduced HP GluN2B levels in males but not in female offspring [127]. In another study, alcohol exposure (10%) during the gestation and lactation period increased GluN2B expression in the synaptic compartment and caused greater low-frequency stimulation (LFS; 600 pulses)-induced LTD compared to control adult CA1 HP slices [111]. In the same study, alcohol exposure reduced LTP. These findings indicate that alcohol exposure during pregnancy increased the highly Ca2+-permeable subunit (GluN2B), leading to an increased NMDA-dependent LTD and a concomitantly reduced LTP magnitude.
As discussed above, different alcohol exposure paradigms during the gestational or postnatal period altered the expression of both GluN2A and GluN2B subunits and therefore caused changes in the ratio between these two critical subunits of NMDARs. Interestingly, the GluN2A and GluN2B subunits have been shown to control bidirectional synaptic plasticity in a given neuronal circuit [61]. Especially when the GluN2A-to-GluN2B ratio is low, LTP is more likely to follow through the studied synapse than LTD and vice versa. In this regard, prenatal alcohol exposure in mice caused a decrease in GluN2B subunits and increased C2-containing GluN1 and GluN3A subunits at the HP DG synapse [30]. It should be noted that GluN1 and GluN3 subunits are poorly permeable to Ca2+, whereas GluN2B is highly permeable to Ca2+. Hence, such changes in subunit expression and the ratio will ultimately modify the plasticity in alcohol-exposed synapses. Alcohol exposure from PD4–9 by gavage reduced the GluN2B subunit in the dorsal HP synaptic compartment, thus increasing the GluN2A to GluN2B subunit ratio [128]. Such changes are thought to be critical for trace fear conditioning [128]. Alcohol exposure (two bottled choices) throughout pregnancy increased GluN1 levels in the synaptosomal membrane fraction without altering GluN2A and GluN2B in the HP DG [109], and PSD-95-associated pools of receptor subunits showed no changes in GluN1 or GluN2A but a decrease in GluN2B [109]. These rearrangements of the NMDA subunits (reduced GluN2B) at the synapse in the DG lead to reduced NMDA-dependent LTP and reduced LTD. Interestingly, in another study, alcohol exposure (10%) during the gestation and lactation period increased the GluN2B subunit in the synaptic compartment in the HP CA1 region [111] due to an increase in the highly Ca2+-permeable subunit. These findings suggest that manipulating NMDAR subunit ratios may offer neuroprotection against developmental alcohol-induced neuronal plasticity and may be associated with behavioral abnormalities.

2.2. The Developmental Alcohol Effects on the GABAergic Neurotransmitter System

Inhibition of neuronal plasticity plays a crucial role in regulating neuronal homeostasis, which is the basis of nervous system function. This inhibition is primarily facilitated by the neurotransmitters GABA and glycine, which activate Cl⁻ permeable ion channels, indicating that the strength of inhibition rests on the Cl⁻ gradient across the synaptic membrane [129]. Therefore, the balance between inhibitory neuronal transmission through GABA and excitatory neuronal transmission through glutamate is indispensable for proper neuronal stability and neurologic function. GABA receptors are classified into GABAA and GABAB [130]. GABAA receptors serve as the major inhibitory neurotransmitter system in the mammalian brain. Each isoform comprises five identical subunits encompassing a central chloride ion-selective channel gated by GABA (Figure 2). GABAA receptors localized to the postsynaptic membrane regulate neuronal inhibition that appears in the millisecond time range, and those localized to the extrasynaptic membrane respond to ambient GABA and are responsible for long-term inhibition [130].
Studies have shown that GABAergic system dysfunction [131,132,133,134,135] likely impacts neurobehavioral abnormalities in offspring exposed to alcohol during development (Table 2). Early studies have explored the influence of gestational alcohol exposure on GABA levels in different brain regions of the offspring. In rats, alcohol exposure (a diet containing 6% ethanol) throughout gestation (4 weeks) increased GABA levels in the cerebral tissues [136]. In rats, alcohol exposure (10%) throughout pregnancy and during the lactation period significantly increased GABA levels in the PFC, olfactory bulb (OFB), anterior colliculus (AC) and amygdala (Amy) tissues from PD21 offspring [137]. The same study found decreased GABA levels in the thalamus, pons, cerebellum, and HP [137]. Alcohol exposure in chick embryos on Days E1–E3 increased glutamate decarboxylase (GAD; a GABAergic neuronal marker) in the E8 embryo’s whole brain [138,139]. In rats, alcohol exposure from GD15 to GD18 increased GABAergic responses in adult frontal and somatosensory cortical neurons [140]. In another study, alcohol exposure using a liquid diet (5% ethanol) throughout gestation caused increased sensitivity of GABAA receptor-stimulated Cl- flux in membrane vesicles prepared from different brain regions, suggesting that such a change in the HP might have contributed to synaptic plasticity defects in adult offspring [131]. In another study, alcohol exposure using a liquid diet from GD0–21 in rats reduced the parvalbumin-expressing GABAergic (PV+) interneurons in the medial septum [141] and anterior cingulate cortex [142] of adult rats. In a guinea pig study, alcohol exposure throughout pregnancy reduced the number of GAD+ ve cells in the somatosensory cortex of adult guinea pigs [143]. In monkeys, alcohol exposure during the first six or the entire 24 weeks of gestation (one day/week) reduced the number of GABA+ ve neurons in the somatosensory and motor cortices of adolescent macaques [144]. In mice, alcohol exposure from GD1–14.5 induced premature GABAergic interneuron tangential migration into the cortical anlage in a 14.5-day-old embryo. In the same study, increased GABA levels and GABA sensitivity of migrating interneurons were observed [145]. In mice, alcohol exposure (binge-type; 5% ethanol) for three days from E13.5 and E16.5 increased the density of median ganglionic eminence-derived interneurons in 16.5-day-old embryos [146]. These alcohol effects persistently increase the number of PV+ interneurons in layer V of the mPFC and potentiate GABAA receptor-mediated synaptic transmission in pyramidal neurons [146]. Other studies have also suggested that early alcohol exposure potentiates the depolarizing effects of GABAA receptors in migrating cells and increases neurogenesis in the medial ganglionic eminence [147,148]. In differentiating human pluripotent stem cell-derived neurons, alcohol exposure (50 mM for 50 days) reduced the transcripts related to GABAergic interneuron specification (i.e., NPY, GSX2, SST and DLX1-6) without affecting interneuron numbers [149]. In mice, alcohol exposure throughout the gestation period increased spontaneous inhibitory postsynaptic current (IPSC) amplitude and area in OFC pyramidal neurons [150], impaired behavioral flexibility and altered OFC activity [151]. Taken together, these findings suggest that the equivalent alcohol exposure during the first and second trimesters of human pregnancy impair interneuron proliferation, differentiation, migration, and/or survival, and contribute to the synaptic plasticity, cognitive and social problems seen in adolescents/adults with documented prenatal alcohol exposure.
Alcohol exposure during human pregnancy, equivalent to the third trimester, has also been shown to have detrimental effects on the GABAergic system. In rats, exposure to alcohol vapor between PD2–6 increased the number of calretinin+ ve interneurons and reduced calbindin+ ve interneurons without affecting PV+ ve interneurons in the primary motor and somatosensory cortex in adolescent offspring (P60) [152]. In the same model, these investigators also found reduced PV+ ve interneurons in the dendritic tree of the striatum (ST) in adolescent offspring (P60) [160]. In mice, exposure to alcohol vapor during gestation (GD12–19) and the neonatal period (PD2–9) reduced cerebellar interneuron numbers at PD16 [153]. In a mouse study, alcohol vapor exposure from PD2–9 reduced interneuron numbers in the adult mouse HP [154]. In the same study, alcohol vapor exposure at PD7 enhanced the number of interneurons that also exhibited activated caspase-3 staining, suggesting that these interneurons are programmed to undergo apoptotic neurodegeneration [154].
Similarly, in PD7 mice, alcohol exposure reduced the numbers of PV+ ve interneurons in the adult mPFC [155], as well as in the HP formation and the pyriform cortex [156,157]. Furthermore, in the same PD7 alcohol exposure model, reduced PV+ ve and calretinin+ ve interneurons were observed in the adult neocortex [158]. Although these alterations significantly contribute to GABAergic neurotransmission defects, the underlying mechanisms that ultimately contribute to neuronal plasticity and behavioral abnormalities require future investigation. Intriguingly, a recent study using the PD7 alcohol vapor model found neuronal degeneration in PD7 mice through inhibition of neuronal activity via the reduced NMDA receptors functions rather than potentiation of Cl current flow through GABAA receptors [159]. In the same study, findings indicated that acute alcohol exposure has no presynaptic or postsynaptic effect on GABAA receptor-mediated synaptic transmission at RSC neurons [159]. However, electrophysiological recordings in slices from adolescent animals showed enhanced peak amplitudes, asynchronous activity, total charge, and reduced rise times of optically evoked GABAA receptor-mediated inhibitory postsynaptic currents [159]. Thus, any of these changes could contribute to the behavioral abnormalities found in animal models of FASD. Therefore, studies to link these two events should be addressed in future preclinical studies. Based on the above studies, it is clear that developmental alcohol exposure has a variable effect on the GABAergic system depending on the brain region and developmental stage at which analysis was made in addition to the alcohol dose used. Increased GABA interneurons may be attributed to enhanced tangential migration [145]. Tangential migration is controlled mainly by GABA signaling. Therefore, reducing ambient GABA could cause defective GABA interneurons migration, which expresses GABA receptors, a target of alcohol. In addition, alcohol has been shown to potentiate GABA signaling by increasing GABA release and receptor signaling leading to premature migration. In addition to these mechanisms, abundance and differences in GABA interneurons composition in each brain region, chloride efflux, growth factors, and calcium signaling defects can contribute to variable GABA interneuronopathy in a brain region-specific manner, which could account for the synaptic and behavioral abnormalities found in FASD and warrant future investigations in this line of research.

2.3. The Developmental Effects of Alcohol on Long-Term Synaptic Plasticity

The current literature indicates that the impact of early alcohol exposure on persistent HP long-term synaptic plasticity (Table 3) is dependent on the alcohol administration paradigm, the developmental period of alcohol exposure, and plasticity induction protocols. Initial observations indicated significantly reduced LTP magnitude in CA1 of young adult male offspring exposed to alcohol during GD1–22 [161]. After this initial study, several investigators examined LTP in the CA1 region using different alcohol exposure paradigms. They reported predominantly reduced LTP in males (See Table 3). However, female offspring exhibited mixed results. For example, alcohol exposure from GD1–22 reduced CA1 LTP in young males and increased it in female offspring [162]. Similar results were observed in another study where alcohol was exposed from GD0–22, and CA1 LTP was performed in male and female young animals [163]. In another study, alcohol exposure from GD1–22 reduced CA1 LTP in male adolescents but not in females [164]. Although the mechanisms responsible for these sex differences are unknown, developmental alcohol-induced changes in the activity of the hypothalamic-pituitary-gonadal axis [165,166,167] may contribute to these dimorphic defects, and future studies to address this are warranted.
A few studies have examined the impact of developmental alcohol on CA1 LTD (Table 3). Reduced LTD in the adolescent CA1 was reported in P7 mice exposed to acute alcohol [169]. However, alcohol exposure throughout pregnancy and lactation (three trimesters) enhanced LTD in CA1 from PD50 offspring [111]. Interestingly, the LTD protocol that failed to induce LTD in controls was able to induce LTD in alcohol-exposed offspring [111]. Lower-dose alcohol exposure limited to only the first two trimesters of gestation failed to affect CA1 LTD in vivo in adolescent male and female alcohol-exposed offspring [180]. In another study, alcohol exposure during GD0–22 reduced LTD in CA1 of male alcohol-exposed offspring. However, it enhanced LTD in females compared to controls [163]. Finally, in a recent study, alcohol exposure was restricted to GD8 and 12 and reduced LTD in adult offspring [172]. Although additional studies are required to explain these sex differences in LTD, these limited findings emphasize that developmental alcohol-induced synaptic plasticity imbalance appears sexually dimorphic.
In the DG, developmental alcohol exposure consistently reduced LTP in male offspring. However, in females, no change or facilitation of LTP was observed (Table 3) in different alcohol exposure paradigms. The DG synaptic plasticity deficits appear to be dependent on the developmental period of alcohol exposure. For example, reduced LTP in male offspring was observed only after the second trimester-equivalent (GD11–21), but not in first- or third-trimester-equivalent exposure [174]. However, in other studies, alcohol exposure from GD1–22 and during the third-trimester equivalent reduced LTP in adolescent offspring [126,178,179]. Few studies have explored LTD in the HP DG region and produced both reduced and enhanced LTD [177,178]. Long-term synaptic plasticity relies on signaling cascades that ultimately control the gene expression that regulates these synaptic activities. Therefore, future studies evaluating the developmental alcohol-induced signaling events and transcription mechanisms mediated by several modulators of glutamatergic and GABAergic neurotransmitter systems can provide a potential relationship between these events.

2.4. The Influence of Developmental Alcohol on Modulators of Glutamatergic and GABAergic Neurotransmitter System

Dopamine (DA) has been shown to modulate excitatory and inhibitory neurotransmission through second messenger signaling systems activated upon receptor stimulation and through receptor cross-talk. Based on efferent projections, three main dopaminergic tracts, the mesostriatal, mesocortical, and mesolimbic, have been reported in the mammalian brain [181]. Furthermore, the mesostriatal system predominantly functions in voluntary movement, the mesocortical system is involved in motivation, attention, and behavior, and the mesolimbic system has been shown to regulate emotion and memory [181]. Interestingly, dopaminergic projections make synaptic connections with GABAergic neurons in the nucleus accumbens (NAc), PFC, and ST. Likewise, dopaminergic neurons also make synaptic connections with glutamatergic neurons in the PFC, ST, and NAc [181]. DA hypofunction has been observed in different animal models (Table 4), suggesting that the dopaminergic system is sensitive to developmental alcohol in certain brain regions [182,183,184,185,186,187,188]. In a rat study, alcohol exposure from GD6–20 caused supersensitivity of somatodendritic DA autoreceptors in the ventral tegmental area (VTA), as measured by extracellular recording [189]. In a similar study, alcohol exposure from GD8 until parturition produced a long-lasting reduction in DA receptor function distinct from the somatodendritic impulse-regulating D2 autoreceptors [190]. In another rat study, alcohol exposure during GD8–20 reduced the number of spontaneously active VTA DA neurons without altering the firing rate or firing pattern [191]. Later, it was found that impaired postnatal development contributed to a persistent reduction in the spontaneous electrical activity of VTA DA neurons in adult animals [192]. The reduction in the spontaneous electrical activity of VTA DA neurons in alcohol-exposed adult offspring was due to altered inputs to VTA DA neurons [193]. In rats, alcohol exposure from GD8–20 enhanced GluA3 subunits in VTA DA neuron (2–12-week-old) dendrites and increased excitatory synaptic strength and the induction of Ca2+-permeable AMPA receptor-dependent LTP, an anti-Hebbian form of LTP [79]. Similar alcohol exposure in rats (GD7–GD20) enhanced basal synaptic transmission [194] and increased presynaptic glutamate release via enhanced D1 receptor function in the corticostriatal pathway in PD30 offspring [194]. These findings suggest that the imbalance between the function of D1 and D2 receptors, resulting from the upregulation of D1 receptors and downregulation of D2 receptors, promoted LTP instead of LTD in alcohol-exposed offspring [194]. Additionally, DA has been shown to bidirectionally regulate GABAA receptor-dependent synaptic transmission by increasing local interneuron excitability through D1 receptors [195] and decreasing quantal GABA release onto pyramidal neurons through a DA type-3 receptor (D3R) [196]. Accordingly, in the rat model, alcohol vapor exposure during PD2–12 decreased DA levels, attenuated D1 receptor-mediated potentiation of sIPSCs and impaired D3R-mediated suppression of mIPSCs in adolescent offspring (PD40–50) pyramidal neurons [197] without affecting D1 and D3 receptor expression. These observations suggest that alcohol exposure during the first and second trimesters of human pregnancy’s equivalent developmental period influences neuronal plasticity via the DA system in adult offspring.
CB1 receptors are expressed at axon terminals of glutamatergic and GABAergic neuron types throughout the brain, and function by modulating many physiological processes, such as synaptic plasticity, cognitive functions, and affective behavior [207,208]. Postsynaptic release of eCBs on demand and binding to presynaptic CB1 receptors suppresses glutamatergic or GABAergic transmission, a process termed eCB acting as a retrograde neurotransmitter [209,210]. The influence of alcohol on eCB/CB1 receptors and the role of eCBs and CB1 receptors in AUDsin adult animals have been extensively studied (for references, see [211]). However, although CB1 regulates glutamatergic and GABAergic function [212], few studies have explored the influence of eCB/CB1 on neuronal survival, plasticity and cognitive function in response to developmental alcohol exposure (Table 4) [43,45,74,198].
Studies have demonstrated that PD7 alcohol exposure causes apoptosis, increases the abundance of eCBs, such as anandamide (AEA), and increases CB1 expression in exposed PD8 mice [45]. Additionally, PD7 alcohol exposure induced apoptosis in PD8 mice, which was absent in mice treated with a CB1 antagonist (SR141716A, SR) before PD7 alcohol exposure and in PD8 global CB1-KO mice [45]. Consistently, CB1 activation during the early developmental period via exposure to cannabinoids at doses similar to those observed in cannabis users causes delays in the maturation of neurotransmitter systems [213]. These delays result in cognitive defects [214] identical to those found in several specific human developmental disorders [215], including FASD [216]. Moreover, the acute administration of Δ9-tetrahydrocannabinol (THC) (the main active constituent of marijuana), which activates CB1 receptors, to P7 rats enhanced the proapoptotic effects of alcohol [217]. Intriguingly, THC coadministered with low-dose alcohol in PD7 rats increased CB1 expression in a brain-region-specific manner in PD8 rat brains [217]. Consistent with these observations, a CB1-KO genotype reduces the susceptibility of PD7 mice to the neurotoxic effects of low-dose [217] or high-dose alcohol [45]. In addition to these immediate effects of alcohol, we and others have demonstrated persistent neurobehavioral defects in adult mice exposed to PD7 alcohol [43,44,45,156,171,218,219,220]. Neither CB1 antagonist-treated PD7 mice nor global CB1-KO mice exhibit LTP, spatial memory or spatial recognition memory abnormalities [44,45] as adults. Consistent with these exciting studies, another study using a zebrafish model suggests that activation of CB1 using the agonist arachidonyl-2-chloroethylamide resulted in a FASD-like phenotype that was reversible using SR [221].
Additionally, cannabinoid administration exacerbated the teratogenic effects of alcohol in embryonic mice and zebrafish models, and CB1 receptor antagonists attenuated these effects [222]. Furthermore, gestational alcohol exposure from the first- to the third-trimester equivalents of human pregnancy in mice resulted in tonically active CB1 in DLS projection neurons [199]. Pharmacologically increasing eCB tone in these studies mimicked the effects of alcohol on synaptic transmission and cognition [199]. In another study, alcohol exposure during GD8–20 resulted in tonic eCB signaling in ST neurons [198]. In the same alcohol exposure paradigm, moderate and high levels of alcohol exposure persistently reduced LFS-induced eCB-mediated LTD in VTA DA neurons in young animals (4–10 weeks) [198]. Alcohol vapor exposure in mice caused increased excitability in DLS medium spiny neurons (MSNs) followed by increased eCB tone [199]. In contrast to the above studies, alcohol exposure between GD8–20 impaired tonic eCB signaling and caused anxiety-like behavior in adult rats [200]. These initial findings collectively indicate that alcohol alters eCB/CB1 function. However, this effect depends on the exposure paradigm (pre- or postnatal exposure) and the brain regions studied. Altered CB1 signaling may influence lasting synaptic, cognitive, and sociobehavioral defects. However, the underlying neural basis through which CB1 contributes to these persistent neuronal and behavioral impairments is unknown and warrants future investigation.
The serotonin neurotransmitter, also called 5-hydroxytryptamine (5-HT), has been shown to bind to 5-HT-gated ion channels (5-HT3 receptors) or G protein-coupled receptors (5-HT1, 5-HT2, and 5-HT4–7 receptors). The majority of 5-HT neurons are located in the raphe nuclei in the brain stem. Some of the 5-HT neurons project to the spinal cord, and others project to the cortex, HP, and hypothalamus (for a review, see [223]). Glutamatergic [224] and GABAergic [225] neurons provide input to 5-HT neurons and modulate their functions in many brain regions. In addition, 5-HT neurons are found as early as E12 in rodents and are thought to influence cell proliferation, differentiation, migration, and synapse formation [226]. Therefore, the function of the 5-HT system in developmental alcohol exposure has been investigated by many (Table 4), as changes in the 5-HT system could have a vast effect on neuronal circuit maturation and behavioral outcomes. In a rat study, alcohol exposure throughout gestation reduced 5-HT, 5-hydroxyindoleacetic acid (5-HIAA), and 5-HT1 receptors in the motor and somatosensory cortex of PD19 offspring [184,201]. A similar reduction in 5-HT immunoreactivity was found in the medial forebrain (MFB) of E15 and 18 embryos exposed to alcohol from GD8–15 [202,203,204]. Another model of gestational alcohol showed reduced 5-HT and its synthesis enzyme, tryptophan hydroxylase (TPH), in the dorsal raphe of 5-week-old offspring [205]. Alcohol exposure throughout the gestation and lactation period reduced 5-HT in the amygdala and cingulate cortex (CC) of offspring (PD77–84), as the 5-HT1 receptor was reduced in the CC region and increased in the Amy region [206]. Alcohol exposure between GD8–20 potentiated glutamate synapses of dorsal raphe nucleus-5-HT neurons and caused anxiety-like behavior in adult rats [200]. These findings indicate that developmental alcohol significantly impairs the 5-HT system. However, the mechanisms are less clear and future investigations on the underlying detrimental effects of developmental alcohol on the 5-HT system and its influence on synaptic plasticity and behavioral outcomes are warranted.

3. Conclusions

Investigations have increased significantly over the past decade into the influence of developmental alcohol exposure on neuronal plasticity mediated by glutamatergic and GABAergic systems and modulators of these two critical neurotransmitters. These studies were made possible thanks to technical and conceptual advances in the neuroscience field over the past decade. Studies have clearly revealed that these neurotransmitter systems in the developing brain are highly vulnerable to alcohol exposure compared to the adult brain. This is because components of developing neurotransmitter systems have unique features that make them specifically sensitive to the harmful effects of low- to high-dose alcohol exposure. A great deal of literature (Table 1, Table 2, Table 3 and Table 4) has demonstrated that developmental alcohol exposure causes defects in the formation and refinement of neuronal circuits/assemblies mediated by glutamatergic, GABAergic, and their modulator signaling events (Figure 3), which are likely to contribute, in part, to the long-lasting structural and functional brain impairments that feature FASD. Neuronal plasticity defects are eventually responsible for the behavioral and cognitive outcomes observed in children with FASD and their increased tendency to exhibit comorbid neuropsychiatric and neurological disorders. In future studies, the underlying mechanisms by which developmental alcohol alters glutamatergic and GABAergic systems should continue to be evaluated, and these studies should be expanded to other modulators of neurotransmitter systems (DA, eCB/CB1, and 5-HT) which may provide the therapeutic potential to treat FASD.

Author Contributions

Conceptualization, B.S.B.; funding acquisition, B.S.B.; Writing-review and editing, B.S.B.; literature and table preparation, B.S.B. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

BSB is supported by NIH/NIAAA grant (# AA019443).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Alc, alcohol; FB, forebrain; SD, Sprague–Dawley; LE, Long-Evans; CE, cortex; WB, Whole brain; i.p, intraperitonial injection; GD, gestational day; PD, postnatal day; NC, neocortex; VTA, ventral tegmental area; mEPSCs, miniature excitatory postsynaptic currents; EPSCs, excitatory postsynaptic currents; mPFC, medial prefrontal cortex; WH, Wistar Han outbred rats; dHP, dorsal HP; A, albino rats; THA, thalmus; CBL, cerebellum, FC, frontal cortex; OLB, olfactory bulbs; Amy, Amygdala; AC, anterior colliculus; GAD, glutamate decarboxylase; mFC, medial frontal cortex; PV, parvalbumin; IN, interneurons; MS, medial septum; ACC, anterior cingulate cortex; SC, somatosensory cortex; MC, motor cortex; PN, pyramidal neurons; OFC, orbitofrontal cortex; IPSCs, inhibitory post-synaptic currents; PYN, pyramidal neurons; CR, calretinin; CB, calbindin; ST, striatum; PC, pyriform cortex; NC, neocortex; RSC, retrosplenial cortex; PPF, paired-pulse potentiation; LTP, long-term potentiation; LTD, long-term depression; HP, Hippocampus; NSDDA, nigrostriatal dopaminergic neurons; DA, dopamine; DL-ST, dorsolateral striatum; D1R, dopamine receptor 1; BLA, basolateral amygdala; D3R, dopamine receptor 3; NAPLE-PLD, N-Arachidonoyl phosphatidylethanolamine–phospholipase D; GDE1, glycerophosphodiesterase (GDE1); CB1, cannabinoid receptor type 1; eCB, endocannabinoids, DRn5-HT, dorsal raphe nucleus-serotonin neurons; DLS, dorsolateral striatum; MSNs, medium spiny neurons; MC, motor cortex; 5-HT, serotonin; 5-HIAA, 5-hydroxy indole acetic acid; SSC, somatosensory cortex; MFB, medial forebrain; MR, median raphe; DR, dorsal raphe; TPH, tryptophan hydroxylase; Amy, amygdala; CC, Cingular cortex, [3H]tritiated thymidine, MK801, Inhibitor of the N-Methyl-D-aspartate; CNS, central nervous system; FASD, Fetal alcohol spectrum disorder; MD, Medial Forebrain; FAS, fetal alcohol syndrome; AMPAR, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; GABA, gamma-aminobutyric acid; NMDAR, N-methyl-D-aspartate receptor; mGluRs, metabotropic glutamate receptors; iGluRs, ionotropic glutamate receptors; GluD, glutamate delta; PD, postnatal day; fEPSPs, field excitatory postsynaptic potentials; AEA, N-arachidonoylethanolamide or anandamide; 2-AG 2-arachidonylglycerol; eCBs, endocannabinoids; CB1; cannabinoid receptor type 1; VGCCs, N-type voltage-gated Ca2+ channels, FMI-43; KCl, potassium chloride; FMI-43, [N-(-3-trethylammoniumpropyl)-4-(-(dibutylamino) styryl) pyridinium dibromide; SR141716A, N-(Piperidin-1-yl)-5-(4-chlorophenyl)-1-(2,4-dichlorophenyl)-4-methyl-1H-pyrazole-3-carboxamide hydrochloride; GD, gestational day; GluR, glutamate receptor subunit; DG, dentate gyrus; Glu N, NMDA receptor subunit; mRNA, messenger ribonucleic acid; PSD95, Postsynaptic density protein 95; CREB, cAMP response element-binding protein; ERK, extracellular signal-regulated kinase; eEPSCs, evoked field excitatory postsynaptic potentials, LFS, low-frequency stimulation; AC, anterior colliculus; E, embryonic day; IPSC, inhibitory postsynaptic current; GSX2 (GS Homeobox 2); NPY, Neuropeptide Y; SST, Somatostatin; DLX, Distal-Less Homeobox, WS, Wistar; A, Albino; AUD, alcohol use disorder.

References

  1. SAMHSA. Center for Behavioral Health Statistics and Quality. 2019 National Survey on Drug Use and Health. Table 2.18B—Alcohol Use in Past Year among Persons Aged 12 or Older, by Age Group and Demographic Characteristics: Percentages, 2018 and 2019; SAMHSA: Washington, DC, USA, 2019. [Google Scholar]
  2. Dejong, K.; Olyaei, A.; Lo, J. Alcohol Use in Pregnancy. Clin. Obstet. Gynecol. 2019, 62, 142–155. [Google Scholar] [CrossRef] [PubMed]
  3. Denny, C.H.; Acero, C.; Naimi, T.; Kim, S. Consumption of Alcohol Beverages and Binge Drinking Among Pregnant Women Aged 18–44 Years—United States, 2015–2017. MMWR Morb. Mortal Wkly. Rep. 2019, 68, 365–368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Muggli, E.; Hearps, S.; Halliday, J.; Elliott, E.; Penington, A.; Thompson, D.; Spittle, A.; Forster, D.; Lewis, S.; Anderson, P. A data driven approach to identify trajectories of prenatal alcohol consumption in an Australian population-based cohort of pregnant women. Sci. Rep. 2022, 12, 4353. [Google Scholar] [CrossRef]
  5. Muggli, E.; O’Leary, C.; Donath, S.; Orsini, F.; Forster, D.; Anderson, P.; Lewis, S.; Nagle, C.; Craig, J.; Elliott, E.; et al. “Did you ever drink more?” A detailed description of pregnant women’s drinking patterns. BMC Public Health 2016, 16, 683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Nykjaer, C.; Alwan, N.; Greenwood, D.; Simpson, N.; Hay, A.; White, K.; Cade, J. Maternal alcohol intake prior to and during pregnancy and risk of adverse birth outcomes: Evidence from a British cohort. J. Epidemiol. Community Health 2014, 68, 542–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Tan, C.H.; Denny, C.; Cheal, N.; Sniezek, J.; Kanny, D. Alcohol use and binge drinking among women of childbearing age—United States, 2011–2013. MMWR Morb. Mortal Wkly. Rep. 2015, 64, 1042–1046. [Google Scholar] [CrossRef] [Green Version]
  8. Jones, K.L.; Smith, D.; Ulleland, C.; Streissguth, P. Pattern of malformation in offspring of chronic alcoholic mothers. Lancet 1973, 1, 1267–1271. [Google Scholar] [CrossRef]
  9. Jones, K.L.; Smith, D.W. Recognition of the fetal alcohol syndrome in early infancy. Lancet 1973, 2, 999–1001. [Google Scholar] [CrossRef]
  10. May, P.A.; Keaster, C.; Bozeman, R.; Goodover, J.; Blankenship, J.; Kalberg, W.; Buckley, D.; Brooks, M.; Hasken, J.; Gossage, J.; et al. Prevalence and characteristics of fetal alcohol syndrome and partial fetal alcohol syndrome in a Rocky Mountain Region City. Drug Alcohol Depend. 2015, 155, 118–127. [Google Scholar] [CrossRef] [Green Version]
  11. Alati, R.; Al Mamun, A.; Williams, G.; O’Callaghan, M.; Najman, J.; Bor, W. In utero alcohol exposure and prediction of alcohol disorders in early adulthood: A birth cohort study. Arch. Gen. Psychiatry 2006, 63, 1009–1016. [Google Scholar] [CrossRef]
  12. Autti-Ramo, I.; Fagerlund, A.; Ervalahti, N.; Loimu, L.; Korkman, M.; Hoyme, H. Fetal alcohol spectrum disorders in Finland: Clinical delineation of 77 older children and adolescents. Am. J. Med. Genet. A 2006, 140, 137–143. [Google Scholar] [CrossRef] [PubMed]
  13. Ceccanti, M.; Spagnolo, P.A.; Tarani, L.; Attilia, M.L.; Chessa, L.; Mancinelli, R.; Stegagno, M.; Sasso, G.F.; Romeo, M.; Jones, K.; et al. Clinical delineation of fetal alcohol spectrum disorders (FASD) in Italian children: Comparison and contrast with other racial/ethnic groups and implications for diagnosis and prevention. Neurosci. Biobehav. Rev. 2007, 31, 270–277. [Google Scholar] [CrossRef] [PubMed]
  14. Spohr, H.L.; Willms, J.; Steinhausen, H. Fetal alcohol spectrum disorders in young adulthood. J. Pediatr. 2007, 150, 175–179. [Google Scholar] [CrossRef] [PubMed]
  15. Streissguth, A.P.; Aase, J.; Clarren, S.; Randels, S.; LaDue, R.; Smith, D. Fetal alcohol syndrome in adolescents and adults. JAMA 1991, 265, 1961–1967. [Google Scholar] [CrossRef] [PubMed]
  16. Sokol, R.J.; Delaney-Black, V.; Nordstrom, B. Fetal alcohol spectrum disorder. JAMA 2003, 290, 2996–2999. [Google Scholar] [CrossRef]
  17. May, P.A.; Baete, A.; Russo, J.; Elliott, A.; Blankenship, J.; Kalberg, W.; Buckley, D.; Brooks, M.; Hasken, J.; Abdul-Rahman, O.; et al. Prevalence and characteristics of fetal alcohol spectrum disorders. Pediatrics 2014, 134, 855–866. [Google Scholar] [CrossRef] [Green Version]
  18. May, P.A.; Chambers, C.; Kalberg, W.; Zellner, J.; Feldman, H.; Buckley, D.; Kopald, D.; Hasken, J.; Xu, R.; Honerkamp-Smith, G.; et al. Prevalence of Fetal Alcohol Spectrum Disorders in 4 US Communities. JAMA 2018, 319, 474–482. [Google Scholar] [CrossRef] [Green Version]
  19. Aragon, A.S.; Kalberg, W.; Buckley, D.; Barela-Scott, L.; Tabachnick, B.; May, P. Neuropsychological study of FASD in a sample of American Indian children: Processing simple versus complex information. Alcohol Clin. Exp. Res. 2008, 32, 2136–2148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Green, C.R.; Mihic, A.; Nikkel, S.; Stade, B.; Rasmussen, C.; Munoz, D.; Reynolds, J. Executive function deficits in children with fetal alcohol spectrum disorders (FASD) measured using the Cambridge Neuropsychological Tests Automated Battery (CANTAB). J. Child Psychol. Psychiatry 2009, 50, 688–697. [Google Scholar] [CrossRef]
  21. Mattson, S.N.; Crocker, N.; Nguyen, T. Fetal alcohol spectrum disorders: Neuropsychological and behavioral features. Neuropsychol. Rev. 2011, 21, 81–101. [Google Scholar] [CrossRef]
  22. Mattson, S.N.; Roesch, S.; Fagerlund, A.; Autti-Ramo, I.; Jones, K.; May, P.; Adnams, C.; Konovalova, V.; Riley, E. Toward a neurobehavioral profile of fetal alcohol spectrum disorders. Alcohol Clin. Exp. Res. 2010, 34, 1640–1650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. May, P.A.; Blankenship, J.; Marais, A.; Gossage, J.; Kalberg, W.; Joubert, B.; Cloete, M.; Barnard, R.; De Vries, M.; Hasken, J.; et al. Maternal alcohol consumption producing fetal alcohol spectrum disorders (FASD): Quantity, frequency, and timing of drinking. Drug Alcohol Depend. 2013, 133, 502–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Lebel, C.; Mattson, S.; Riley, E.; Jones, K.; Adnams, C.; May, P.; Bookheimer, S.; O’Connor, M.J.; Narr, K.; Kan, E.; et al. A longitudinal study of the long-term consequences of drinking during pregnancy: Heavy in utero alcohol exposure disrupts the normal processes of brain development. J. Neurosci. 2012, 32, 15243–15251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Norman, A.L.; O’Brien, J.W.; Spadoni, A.; Tapert, S.; Jones, K.; Riley, E.; Mattson, S. A functional magnetic resonance imaging study of spatial working memory in children with prenatal alcohol exposure: Contribution of familial history of alcohol use disorders. Alcohol Clin. Exp. Res. 2013, 37, 132–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Sood, B.; Delaney-Black, V.; Covington, C.; Nordstrom-Klee, B.; Ager, J.; Templin, T.; Janisse, J.; Martier, S.; Sokol, R. Prenatal alcohol exposure and childhood behavior at age 6 to 7 years: I. dose-response effect. Pediatrics 2001, 108, E34. [Google Scholar] [CrossRef] [Green Version]
  27. Streissguth, A.P.; Bookstein, F.; Sampson, P.; Barr, H. Neurobehavioral effects of prenatal alcohol: Part III. PLS analyses of neuropsychologic tests. Neurotoxicol. Teratol. 1989, 11, 493–507. [Google Scholar] [CrossRef] [Green Version]
  28. Guerri, C.; Bazinet, A.; Riley, E. Foetal Alcohol Spectrum Disorders and alterations in brain and behaviour. Alcohol Alcohol 2009, 44, 108–114. [Google Scholar] [CrossRef] [Green Version]
  29. Basavarajappa, B.S. Fetal Alcohol Spectrum Disorder: Potential Role of Endocannabinoids Signaling. Brain Sci. 2015, 5, 456–493. [Google Scholar] [CrossRef] [Green Version]
  30. Brady, M.L.; Diaz, M.; Iuso, A.; Everett, J.; Valenzuela, C.; Caldwell, K. Moderate prenatal alcohol exposure reduces plasticity and alters NMDA receptor subunit composition in the dentate gyrus. J. Neurosci. 2013, 33, 1062–1067. [Google Scholar] [CrossRef] [Green Version]
  31. Brolese, G.; Lunardi, P.; Broetto, N.; Engelke, D.; Lirio, F.; Batassini, C.; Tramontina, A.; Goncalves, C. Moderate prenatal alcohol exposure alters behavior and neuroglial parameters in adolescent rats. Behav. Brain Res. 2014, 269, 175–184. [Google Scholar] [CrossRef]
  32. Cui, Z.J.; Zhao, K.; Zhao, H.; Yu, D.; Niu, Y.; Zhang, J.; Deng, J. Prenatal alcohol exposure induces long-term changes in dendritic spines and synapses in the mouse visual cortex. Alcohol Alcohol 2010, 45, 312–319. [Google Scholar] [CrossRef] [PubMed]
  33. Downing, C.; Johnson, T.; Larson, C.; Leakey, T.; Siegfried, R.; Rafferty, T.; Cooney, C. Subtle decreases in DNA methylation and gene expression at the mouse Igf2 locus following prenatal alcohol exposure: Effects of a methyl-supplemented diet. Alcohol 2011, 45, 65–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Goggin, S.L.; Caldwell, K.; Cunningham, L.; Allan, A. Prenatal alcohol exposure alters p35, CDK5 and GSK3beta in the medial frontal cortex and hippocampus of adolescent mice. Toxicol. Rep. 2014, 1, 544–553. [Google Scholar] [CrossRef] [Green Version]
  35. Goodlett, C.R.; Horn, K.H. Mechanisms of alcohol-induced damage to the developing nervous system. Alcohol Res. Health 2001, 25, 175–184. [Google Scholar]
  36. Lunde-Young, R.; Ramirez, J.; Naik, V.; Orzabal, M.; Lee, J.; Konganti, K.; Hillhouse, A.; Threadgill, D.; Ramadoss, J. Hippocampal transcriptome reveals novel targets of FASD pathogenesis. Brain Behav. 2019, 9, e01334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Margret, C.P.; Li, C.; Chappell, T.; Elberger, A.; Matta, S.; Waters, R. Prenatal alcohol exposure delays the development of the cortical barrel field in neonatal rats. Exp. Brain Res. 2006, 172, 1–13. [Google Scholar] [CrossRef]
  38. Nagre, N.N.; Subbanna, S.; Shivakumar, M.; Psychoyos, D.; Basavarajappa, B. CB1-receptor knockout neonatal mice are protected against ethanol-induced impairments of DNMT1, DNMT3A, and DNA methylation. J. Neurochem. 2015, 132, 429–442. [Google Scholar] [CrossRef] [Green Version]
  39. Ozer, E.; Sarioglu, S.; Gure, A. Effects of prenatal ethanol exposure on neuronal migration, neuronogenesis and brain myelination in the mice brain. Clin. Neuropathol. 2000, 19, 21–25. [Google Scholar]
  40. Sari, Y. Activity-dependent neuroprotective protein-derived peptide, NAP, preventing alcohol-induced apoptosis in fetal brain of C57BL/6 mouse. Neuroscience 2009, 158, 1426–1435. [Google Scholar] [CrossRef] [Green Version]
  41. Schaffner, S.L.; Lussier, A.; Baker, J.; Goldowitz, D.; Hamre, K.; Kobor, M. Neonatal Alcohol Exposure in Mice Induces Select Differentiation- and Apoptosis-Related Chromatin Changes Both Independent of and Dependent on Sex. Front. Genet. 2020, 11, 35. [Google Scholar] [CrossRef]
  42. Siler-Marsiglio, K.I.; Madorsky, I.; Pan, Q.; Paiva, M.; Neeley, A.; Shaw, G.; Heaton, M. Effects of acute ethanol exposure on regulatory mechanisms of Bcl-2-associated apoptosis promoter, bad, in neonatal rat cerebellum: Differential effects during vulnerable and resistant developmental periods. Alcohol Clin. Exp. Res. 2006, 30, 1031–1038. [Google Scholar] [CrossRef] [PubMed]
  43. Subbanna, S.; Nagre, N.; Shivakumar, M.; Joshi, V.; Psychoyos, D.; Kutlar, A.; Umapathy, N.; Basavarajappa, B. CB1R-Mediated Activation of Caspase-3 Causes Epigenetic and Neurobehavioral Abnormalities in Postnatal Ethanol-Exposed Mice. Front. Mol. Neurosci. 2018, 11, 45. [Google Scholar] [CrossRef] [PubMed]
  44. Subbanna, S.; Nagre, N.; Umapathy, N.; Pace, B.; Basavarajappa, B. Ethanol exposure induces neonatal neurodegeneration by enhancing CB1R Exon1 histone H4K8 acetylation and up-regulating CB1R function causing neurobehavioral abnormalities in adult mice. Int. J. Neuropsychopharmacol. 2015, 18, 1–15. [Google Scholar] [CrossRef] [Green Version]
  45. Subbanna, S.; Shivakumar, M.; Psychoyos, D.; Xie, S.; Basavarajappa, B. Anandamide-CB1 Receptor Signaling Contributes to Postnatal Ethanol-Induced Neonatal Neurodegeneration, Adult Synaptic and Memory Deficits. J. Neuosci. 2013, 33, 6350–6366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Subbanna, S.; Shivakumar, M.; Umapathy, N.; Saito, M.; Mohan, P.; Kumar, A.; Nixonc, R.; Verin, A.; Psychoyos, D.; Basavarajappa, B. G9a-Mediated Histone Methylation Regulates Ethanol-Induced Neurodegeneration in the Neonatal Mouse Brain. Neurobiol. Dis. 2013, 54, 475–485. [Google Scholar] [CrossRef] [Green Version]
  47. Toso, L.; Roberson, R.; Woodard, J.; Abebe, D.; Spong, C. Prenatal alcohol exposure alters GABA(A)alpha5 expression: A mechanism of alcohol-induced learning dysfunction. Am. J. Obstet. Gynecol. 2006, 195, 522–527. [Google Scholar] [CrossRef]
  48. Zhang, W.; Peng, C.; Zheng, M.; Gao, W.; Zhu, J.; Lv, T.; Liu, L.; Liu, Z.; Li, H.; Xv, Y.; et al. Prenatal alcohol exposure causes the over-expression of DHAND and EHAND by increasing histone H3K14 acetylation in C57 BL/6 mice. Toxicol. Lett. 2014, 228, 140–146. [Google Scholar] [CrossRef] [PubMed]
  49. Appelbaum, L.G.; Shenasa, M.; Stolz, L.; Daskalakis, Z. Synaptic plasticity and mental health: Methods, challenges and opportunities. Neuropsychopharmacology 2022, 48, 113–120. [Google Scholar] [CrossRef]
  50. Caporale, N.; Dan, Y. Spike timing-dependent plasticity: A Hebbian learning rule. Annu. Rev. Neurosci. 2008, 31, 25–46. [Google Scholar] [CrossRef] [Green Version]
  51. Markram, H.; Gerstner, W.; Sjostrom, P. A history of spike-timing-dependent plasticity. Front. Synaptic. Neurosci. 2011, 3, 4. [Google Scholar] [CrossRef] [Green Version]
  52. Markram, H.; Lubke, J.; Frotscher, M.; Sakmann, B. Regulation of synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science 1997, 275, 213–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Froemke, R.C.; Dan, Y. Spike-timing-dependent synaptic modification induced by natural spike trains. Nature 2002, 416, 433–438. [Google Scholar] [CrossRef] [PubMed]
  54. Malenka, R.C.; Bear, M.F. LTP and LTD: An embarrassment of riches. Neuron 2004, 44, 5–21. [Google Scholar] [CrossRef] [Green Version]
  55. Chater, T.E.; Goda, Y. The role of AMPA receptors in postsynaptic mechanisms of synaptic plasticity. Front. Cell Neurosci. 2014, 8, 401. [Google Scholar] [CrossRef] [PubMed]
  56. Hollmann, M.; Heinemann, S. Cloned glutamate receptors. Annu. Rev. Neurosci. 1994, 17, 31–108. [Google Scholar] [CrossRef] [PubMed]
  57. Seeburg, P.H. The TiPS/TINS lecture: The molecular biology of mammalian glutamate receptor channels. Trends Pharmacol. Sci. 1993, 14, 297–303. [Google Scholar] [CrossRef] [PubMed]
  58. Dore, K.; Aow, J.; Malinow, R. The Emergence of NMDA Receptor Metabotropic Function: Insights from Imaging. Front. Synaptic. Neurosci. 2016, 8, 20. [Google Scholar] [CrossRef] [Green Version]
  59. Valbuena, S.; Lerma, J. Non-canonical Signaling, the Hidden Life of Ligand-Gated Ion Channels. Neuron 2016, 92, 316–329. [Google Scholar] [CrossRef] [Green Version]
  60. Lau, C.G.; Zukin, R.S. NMDA receptor trafficking in synaptic plasticity and neuropsychiatric disorders. Nat. Rev. Neurosci. 2007, 8, 413–426. [Google Scholar] [CrossRef]
  61. Paoletti, P.; Bellone, C.; Zhou, Q. NMDA receptor subunit diversity: Impact on receptor properties, synaptic plasticity and disease. Nat. Rev. Neurosci. 2013, 14, 383–400. [Google Scholar] [CrossRef]
  62. Parsons, M.P.; Raymond, L.A. Extrasynaptic NMDA receptor involvement in central nervous system disorders. Neuron 2014, 82, 279–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Traynelis, S.F.; Wollmuth, L.; McBain, C.; Menniti, F.; Vance, K.; Ogden, K.; Hansen, K.; Yuan, H.; Myers, S.; Dingledine, R. Glutamate receptor ion channels: Structure, regulation, and function. Pharmacol. Rev. 2010, 62, 405–496. [Google Scholar] [CrossRef]
  64. Dildy-Mayfield, J.E.; Harris, R.A. Comparison of ethanol sensitivity of rat brain kainate, DL-alpha-amino-3-hydroxy-5-methyl-4-isoxalone proprionic acid and N-methyl-D-aspartate receptors expressed in Xenopus oocytes. J. Pharmacol. Exp. Ther. 1992, 262, 487–494. [Google Scholar] [PubMed]
  65. Wirkner, K.; Eberts, C.; Poelchen, W.; Allgaier, C.; Illes, P. Mechanism of inhibition by ethanol of NMDA and AMPA receptor channel functions in cultured rat cortical neurons. Naunyn Schmiedebergs Arch. Pharmacol. 2000, 362, 568–576. [Google Scholar] [PubMed]
  66. Moykkynen, T.; Korpi, E.; Lovinger, D. Ethanol inhibits alpha-amino-3-hydyroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor function in central nervous system neurons by stabilizing desensitization. J. Pharmacol. Exp. Ther. 2003, 306, 546–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Mameli, M.; Zamudio, P.; Carta, M.; Valenzuela, C. Developmentally regulated actions of alcohol on hippocampal glutamatergic transmission. J. Neurosci. 2005, 25, 8027–8036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Puglia, M.P.; Valenzuela, C.F. Ethanol acutely inhibits ionotropic glutamate receptor-mediated responses and long-term potentiation in the developing CA1 hippocampus. Alcohol Clin. Exp. Res. 2010, 34, 594–606. [Google Scholar] [CrossRef] [Green Version]
  69. Durand, G.M.; Kovalchuk, Y.; Konnerth, A. Long-term potentiation and functional synapse induction in developing hippocampus. Nature 1996, 381, 71–75. [Google Scholar] [CrossRef]
  70. Molnar, E.; Pickard, L.; Duckworth, J. Developmental changes in ionotropic glutamate receptors: Lessons from hippocampal synapses. Neuroscientist 2002, 8, 143–153. [Google Scholar] [CrossRef]
  71. Basavarajappa, B.S.; Ninan, I.; Arancio, O. Acute Ethanol Suppresses Glutamatergic Neurotransmission through Endocannabinoids in Hippocampal Neurons. J. Neurochem. 2008, 107, 1001–1013. [Google Scholar] [CrossRef] [Green Version]
  72. Twitchell, W.; Brown, S.; Mackie, K. Cannabinoids inhibit N- and P/Q-type calcium channels in cultured rat hippocampal neurons. J. Neurophysiol. 1997, 78, 43–50. [Google Scholar] [CrossRef] [PubMed]
  73. Maldve, R.E.; Chen, X.; Zhang, T.; Morrisett, R. Ethanol selectively inhibits enhanced vesicular release at excitatory synapses: Real-time visualization in intact hippocampal slices. Alcohol Clin. Exp. Res. 2004, 28, 143–152. [Google Scholar] [CrossRef] [PubMed]
  74. Joshi, V.; Subbanna, S.; Shivakumar, M.; Basavarajappa, B. CB1R regulates CDK5 signaling and epigenetically controls Rac1 expression contributing to neurobehavioral abnormalities in mice postnatally exposed to ethanol. Neuropsychopharmacology 2019, 44, 514–525. [Google Scholar] [CrossRef] [Green Version]
  75. Shivakumar, M.; Subbanna, S.; Joshi, V.; Basavarajappa, B. Postnatal Ethanol Exposure Activates HDAC-Mediated Histone Deacetylation, Impairs Synaptic Plasticity Gene Expression and Behavior in Mice. Int. J. Neuropsychopharmacol. 2020, 23, 324–338. [Google Scholar] [CrossRef] [PubMed]
  76. Bellinger, F.P.; Davidson, M.; Bedi, K.; Wilce, P. Neonatal ethanol exposure reduces AMPA but not NMDA receptor levels in the rat neocortex. Brain Res. Dev. Brain Res. 2002, 136, 77–84. [Google Scholar] [CrossRef] [PubMed]
  77. Dettmer, T.S.; Barnes, A.; Iqbal, U.; Bailey, C.; Reynolds, J.; Brien, J.; Valenzuela, C. Chronic prenatal ethanol exposure alters ionotropic glutamate receptor subunit protein levels in the adult guinea pig cerebral cortex. Alcohol Clin. Exp. Res. 2003, 27, 677–681. [Google Scholar] [CrossRef] [PubMed]
  78. Vaglenova, J.; Pandiella, N.; Wijayawardhane, N.; Vaithianathan, T.; Birru, S.; Breese, C.; Suppiramaniam, V.; Randal, C. Aniracetam reversed learning and memory deficits following prenatal ethanol exposure by modulating functions of synaptic AMPA receptors. Neuropsychopharmacology 2008, 33, 1071–1083. [Google Scholar] [CrossRef] [Green Version]
  79. Hausknecht, K.; Haj-Dahmane, S.; Shen, Y.; Vezina, P.; Dlugos, C.; Shen, R. Excitatory synaptic function and plasticity is persistently altered in ventral tegmental area dopamine neurons after prenatal ethanol exposure. Neuropsychopharmacology 2015, 40, 893–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Louth, E.L.; Bignell, W.; Taylor, C.; Bailey, C. Developmental Ethanol Exposure Leads to Long-Term Deficits in Attention and Its Underlying Prefrontal Circuitry. eNeuro 2016, 3, 1–21. [Google Scholar] [CrossRef] [Green Version]
  81. Takahashi, Y.; Yamashita, R.; Okano, H.; Takashima, K.; Ogawa, B.; Ojiro, R.; Tang, Q.; Ozawa, S.; Woo, G.; Yoshida, T.; et al. Aberrant neurogenesis and late onset suppression of synaptic plasticity as well as sustained neuroinflammation in the hippocampal dentate gyrus after developmental exposure to ethanol in rats. Toxicology 2021, 462, 152958. [Google Scholar] [CrossRef]
  82. Montagud-Romero, S.; Cantacorps, L.; Fernandez-Gomez, F.; Nunez, C.; Minarro, J.; Rodriguez-Arias, M.; Milanes, M.; Valverde, O. Unraveling the molecular mechanisms involved in alcohol intake and withdrawal in adolescent mice exposed to alcohol during early life stages. Prog. Neuropsychopharmacol. Biol. Psychiatry 2021, 104, 110025. [Google Scholar] [CrossRef]
  83. Liu, S.J.; Savtchouk, I. Ca(2+) permeable AMPA receptors switch allegiances: Mechanisms and consequences. J. Physiol. 2012, 590, 13–20. [Google Scholar] [CrossRef] [PubMed]
  84. Cull-Candy, S.; Brickley, S.; Farrant, M. NMDA receptor subunits: Diversity, development and disease. Curr. Opin. Neurobiol. 2001, 11, 327–335. [Google Scholar] [CrossRef] [PubMed]
  85. Cull-Candy, S.G.; Leszkiewicz, D.N. Role of distinct NMDA receptor subtypes at central synapses. Sci. STKE 2004, 2004, re16. [Google Scholar] [CrossRef] [PubMed]
  86. Hansen, K.B.; Wollmuth, L.; Bowie, D.; Furukawa, H.; Menniti, F.; Sobolevsky, A.; Swanson, G.; Swanger, S.; Greger, I.; Nakagawa, T.; et al. Structure, Function, and Pharmacology of Glutamate Receptor Ion Channels. Pharmacol. Rev. 2021, 73, 298–487. [Google Scholar] [CrossRef] [PubMed]
  87. Blevins, T.; Mirshahi, T.; Chandler, L.; Woodward, J. Effects of acute and chronic ethanol exposure on heteromeric N-methyl-D-aspartate receptors expressed in HEK 293 cells. J. Neurochem. 1997, 69, 2345–2354. [Google Scholar] [CrossRef]
  88. Mirshahi, T.; Woodward, J.J. Ethanol sensitivity of heteromeric NMDA receptors: Effects of subunit assembly, glycine and NMDAR1 Mg(2+)-insensitive mutants. Neuropharmacology 1995, 34, 347–355. [Google Scholar] [CrossRef]
  89. Ariwodola, O.J.; Weiner, J.L. Ethanol potentiation of GABAergic synaptic transmission may be self-limiting: Role of presynaptic GABA(B) receptors. J. Neurosci. 2004, 24, 10679–10686. [Google Scholar] [CrossRef] [Green Version]
  90. Wright, J.M.; Peoples, R.; Weight, F. Single-channel and whole-cell analysis of ethanol inhibition of NMDA-activated currents in cultured mouse cortical and hippocampal neurons. Brain Res. 1996, 738, 249–256. [Google Scholar] [CrossRef]
  91. Xu, M.; Smothers, C.; Trudell, J.; Woodward, J. Ethanol inhibition of constitutively open N-methyl-D-aspartate receptors. J. Pharmacol. Exp. Ther. 2012, 340, 218–226. [Google Scholar] [CrossRef] [Green Version]
  92. Ren, H.; Zhao, Y.; Dwyer, D.; Peoples, R. Interactions among positions in the third and fourth membrane-associated domains at the intersubunit interface of the N-methyl-D-aspartate receptor forming sites of alcohol action. J. Biol. Chem. 2012, 287, 27302–27312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Ronald, K.M.; Mirshahi, T.; Woodward, J. Ethanol inhibition of N-methyl-D-aspartate receptors is reduced by site-directed mutagenesis of a transmembrane domain phenylalanine residue. J. Biol. Chem. 2001, 276, 44729–44735. [Google Scholar] [CrossRef] [PubMed]
  94. Smothers, C.T.; Woodward, J.J. Effects of amino acid substitutions in transmembrane domains of the NR1 subunit on the ethanol inhibition of recombinant N-methyl-D-aspartate receptors. Alcohol Clin. Exp. Res. 2006, 30, 523–530. [Google Scholar] [CrossRef] [PubMed]
  95. Savage, D.D.; Montano, C.; Otero, M.; Paxton, L. Prenatal ethanol exposure decreases hippocampal NMDA-sensitive [3H]-glutamate binding site density in 45-day-old rats. Alcohol 1991, 8, 193–201. [Google Scholar] [CrossRef]
  96. Abdollah, S.; Brien, J.F. Effect of chronic maternal ethanol administration on glutamate and N-methyl-D-aspartate binding sites in the hippocampus of the near-term fetal guinea pig. Alcohol 1995, 12, 377–382. [Google Scholar] [CrossRef]
  97. Diaz-Granados, J.L.; Spuhler-Phillips, K.; Lilliquist, M.; Amsel, A.; Leslie, S. Effects of prenatal and early postnatal ethanol exposure on [3H]MK-801 binding in rat cortex and hippocampus. Alcohol Clin. Exp. Res. 1997, 21, 874–881. [Google Scholar] [CrossRef]
  98. Rema, V.; Ebner, F.F. Effect of enriched environment rearing on impairments in cortical excitability and plasticity after prenatal alcohol exposure. J. Neurosci. 1999, 19, 10993–11006. [Google Scholar] [CrossRef] [Green Version]
  99. Naassila, M.; Daoust, M. Effect of prenatal and postnatal ethanol exposure on the developmental profile of mRNAs encoding NMDA receptor subunits in rat hippocampus. J. Neurochem. 2002, 80, 850–860. [Google Scholar] [CrossRef] [Green Version]
  100. Nixon, K.; Hughes, P.; Amsel, A.; Leslie, S. NMDA receptor subunit expression following early postnatal exposure to ethanol. Brain Res. Dev. Brain Res. 2002, 139, 295–299. [Google Scholar] [CrossRef]
  101. Honse, Y.; Nixon, K.; Browning, M.; Leslie, S. Cell surface expression of NR1 splice variants and NR2 subunits is modified by prenatal ethanol exposure. Neuroscience 2003, 122, 689–698. [Google Scholar] [CrossRef]
  102. Nixon, K.; Hughes, P.; Amsel, A.; Leslie, S. NMDA receptor subunit expression after combined prenatal and postnatal exposure to ethanol. Alcohol Clin. Exp. Res. 2004, 28, 105–112. [Google Scholar] [CrossRef] [PubMed]
  103. Toso, L.; Poggi, S.; Abebe, D.; Roberson, R.; Dunlap, V.; Park, J.; Spong, C. N-methyl-D-aspartate subunit expression during mouse development altered by in utero alcohol exposure. Am. J. Obstet. Gynecol. 2005, 193, 1534–1539. [Google Scholar] [CrossRef] [PubMed]
  104. Iqbal, U.; Brien, J.; Kapoor, A.; Matthews, S.; Reynolds, J. Chronic prenatal ethanol exposure increases glucocorticoid-induced glutamate release in the hippocampus of the near-term foetal guinea pig. J. Neuroendocrinol. 2006, 18, 826–834. [Google Scholar] [CrossRef]
  105. Barbier, E.; Pierrefiche, O.; Vaudry, D.; Vaudry, H.; Daoust, M.; Naassila, M. Long-term alterations in vulnerability to addiction to drugs of abuse and in brain gene expression after early life ethanol exposure. Neuropharmacology 2008, 55, 1199–1211. [Google Scholar] [CrossRef] [PubMed]
  106. Staples, M.C.; Porch, M.; Savage, D. Impact of combined prenatal ethanol and prenatal stress exposures on markers of activity-dependent synaptic plasticity in rat dentate gyrus. Alcohol 2014, 48, 523–532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Hughes, P.D.; Kim, Y.; Randall, P.; Leslie, S. Effect of prenatal ethanol exposure on the developmental profile of the NMDA receptor subunits in rat forebrain and hippocampus. Alcohol Clin. Exp. Res. 1998, 22, 1255–1261. [Google Scholar] [CrossRef] [PubMed]
  108. Sickmann, H.M.; Patten, A.; Morch, K.; Sawchuk, S.; Zhang, C.; Parton, R.; Szlavik, L.; Christie, B. Prenatal ethanol exposure has sex-specific effects on hippocampal long-term potentiation. Hippocampus 2014, 24, 54–64. [Google Scholar] [CrossRef]
  109. Samudio-Ruiz, S.L.; Allan, A.; Sheema, S.; Caldwell, K. Hippocampal N-methyl-D-aspartate receptor subunit expression profiles in a mouse model of prenatal alcohol exposure. Alcohol Clin. Exp. Res. 2010, 34, 342–353. [Google Scholar] [CrossRef] [Green Version]
  110. Bird, C.W.; Candelaria-Cook, F.; Magcalas, C.; Davies, S.; Valenzuela, C.; Savage, D.; Hamilton, D. Moderate prenatal alcohol exposure enhances GluN2B containing NMDA receptor binding and ifenprodil sensitivity in rat agranular insular cortex. PLoS One 2015, 10, e0118721. [Google Scholar] [CrossRef]
  111. Kervern, M.; de Ferron, B.S.; Alaux-Cantin, S.; Fedorenko, O.; Antol, J.; Naassila, M.; Pierrefiche, O. Aberrant NMDA-dependent LTD after perinatal ethanol exposure in young adult rat hippocampus. Hippocampus 2015, 25, 912–923. [Google Scholar] [CrossRef]
  112. Zink, M.; Ferbert, T.; Frank, S.; Seufert, P.; Gebicke-Haerter, P.; Spanagel, R. Perinatal exposure to alcohol disturbs spatial learning and glutamate transmission-related gene expression in the adult hippocampus. Eur. J. Neurosci. 2011, 34, 457–468. [Google Scholar] [CrossRef] [PubMed]
  113. Castellani, G.C.; Quinlan, E.; Bersani, F.; Cooper, L.; Shouval, H. A model of bidirectional synaptic plasticity: From signaling network to channel conductance. Learn Mem. 2005, 12, 423–432. [Google Scholar] [CrossRef] [PubMed]
  114. Hughes, P.D.; Wilson, W.; Leslie, S. Effect of gestational ethanol exposure on the NMDA receptor complex in rat forebrain: From gene transcription to cell surface. Brain Res. Dev. Brain Res. 2001, 129, 135–145. [Google Scholar] [CrossRef]
  115. Krapivinsky, G.; Krapivinsky, L.; Manasian, Y.; Ivanov, A.; Tyzio, R.; Pellegrino, C.; Ben-Ari, Y.; Clapham, D.; Medina, I. The NMDA receptor is coupled to the ERK pathway by a direct interaction between NR2B and RasGRF1. Neuron 2003, 40, 775–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Samudio-Ruiz, S.L.; Allan, A.; Valenzuela, C.; Perrone-Bizzozero, N.; Caldwell, K. Prenatal ethanol exposure persistently impairs NMDA receptor-dependent activation of extracellular signal-regulated kinase in the mouse dentate gyrus. J. Neurochem. 2009, 109, 1311–1323. [Google Scholar] [CrossRef]
  117. Sweatt, J.D. Mitogen-activated protein kinases in synaptic plasticity and memory. Curr. Opin. Neurobiol. 2004, 14, 311–317. [Google Scholar] [CrossRef] [PubMed]
  118. Atkins, C.M.; Selcher, J.; Petraitis, J.; Trzaskos, J.; Sweatt, J. The MAPK cascade is required for mammalian associative learning. Nat. Neurosci. 1998, 1, 602–609. [Google Scholar] [CrossRef] [PubMed]
  119. English, J.D.; Sweatt, J.D. Activation of p42 mitogen-activated protein kinase in hippocampal long term potentiation. J. Biol. Chem. 1996, 271, 24329–24332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Savage, D.D.; Becher, M.; de la Torre, A.; Sutherland, R. Dose-dependent effects of prenatal ethanol exposure on synaptic plasticity and learning in mature offspring. Alcohol Clin. Exp. Res. 2002, 26, 1752–1758. [Google Scholar] [CrossRef]
  121. Sutherland, R.J.; McDonald, R.; Savage, D. Prenatal exposure to moderate levels of ethanol can have long-lasting effects on hippocampal synaptic plasticity in adult offspring. Hippocampus 1997, 7, 232–238. [Google Scholar] [CrossRef]
  122. Allan, A.M.; Chynoweth, J.; Tyler, L.; Caldwell, K. A mouse model of prenatal ethanol exposure using a voluntary drinking paradigm. Alcohol Clin. Exp. Res. 2003, 27, 2009–2016. [Google Scholar] [CrossRef] [PubMed]
  123. Gardoni, F.; Mauceri, D.; Malinverno, M.; Polli, F.; Costa, C.; Tozzi, A.; Siliquini, S.; Picconi, B.; Cattabeni, F.; Calabresi, P.; et al. Decreased NR2B subunit synaptic levels cause impaired long-term potentiation but not long-term depression. J. Neurosci. 2009, 29, 669–677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Titterness, A.K.; Christie, B.R. Prenatal ethanol exposure enhances NMDAR-dependent long-term potentiation in the adolescent female dentate gyrus. Hippocampus 2012, 22, 69–81. [Google Scholar] [CrossRef] [PubMed]
  125. Licheri, V.; Chandrasekaran, J.; Bird, C.; Valenzuela, C.; Brigman, J. Sex-specific effect of prenatal alcohol exposure on N-methyl-D-aspartate receptor function in orbitofrontal cortex pyramidal neurons of mice. Alcohol Clin. Exp. Res. 2021, 45, 1994–2005. [Google Scholar] [CrossRef] [PubMed]
  126. Plaza-Briceno, W.; Estay, S.; de la Fuente-Ortega, E.; Gutierrez, C.; Sanchez, G.; Hidalgo, C.; Chavez, A.; Haeger, P. N-Methyl-d-Aspartate Receptor Modulation by Nicotinamide Adenine Dinucleotide Phosphate Oxidase Type 2 Drives Synaptic Plasticity and Spatial Memory Impairments in Rats Exposed Pre- and Postnatally to Ethanol. Antioxid. Redox. Signal. 2020, 32, 602–617. [Google Scholar] [CrossRef] [PubMed]
  127. Ieraci, A.; Herrera, D.G. Early Postnatal Ethanol Exposure in Mice Induces Sex-Dependent Memory Impairment and Reduction of Hippocampal NMDA-R2B Expression in Adulthood. Neuroscience 2020, 427, 105–115. [Google Scholar] [CrossRef]
  128. Goodfellow, M.J.; Abdulla, K.; Lindquist, D. Neonatal Ethanol Exposure Impairs Trace Fear Conditioning and Alters NMDA Receptor Subunit Expression in Adult Male and Female Rats. Alcohol Clin. Exp. Res. 2016, 40, 309–318. [Google Scholar] [CrossRef]
  129. Pressey, J.C.; de Saint-Rome, M.; Raveendran, V.; Woodin, M. Chloride transporters controlling neuronal excitability. Physiol. Rev. 2022, 103, 1095–1135. [Google Scholar] [CrossRef]
  130. Sigel, E.; Steinmann, M.E. Structure, function, and modulation of GABA(A) receptors. J. Biol. Chem. 2012, 287, 40224–40231. [Google Scholar] [CrossRef] [Green Version]
  131. Allan, A.M.; Wu, H.; Paxton, L.; Savage, D. Prenatal ethanol exposure alters the modulation of the gamma-aminobutyric acidA1 receptor-gated chloride ion channel in adult rat offspring. J. Pharmacol. Exp. Ther. 1998, 284, 250–257. [Google Scholar]
  132. Bailey, C.D.; Brien, J.; Reynolds, J. Chronic prenatal ethanol exposure increases GABA(A) receptor subunit protein expression in the adult guinea pig cerebral cortex. J. Neurosci. 2001, 21, 4381–4389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Hsiao, S.H.; Parrish, A.; Nahm, S.; Abbott, L.; McCool, B.; Frye, G. Effects of early postnatal ethanol intubation on GABAergic synaptic proteins. Brain Res. Dev. Brain Res. 2002, 138, 177–185. [Google Scholar] [CrossRef] [PubMed]
  134. Sebe, J.Y.; Eggers, E.; Berger, A. Differential effects of ethanol on GABA(A) and glycine receptor-mediated synaptic currents in brain stem motoneurons. J. Neurophysiol. 2003, 90, 870–875. [Google Scholar] [CrossRef]
  135. Spear, L.P.; Penson, J.; Linville, D. GABA and behavioral inhibition in the neonatal rat pup. Psychopharmacology 1986, 90, 106–111. [Google Scholar] [CrossRef] [PubMed]
  136. Rawat, A.K. Developmental changes in the brain levels of neurotransmitters as influenced by maternal ethanol consumption in the rat. J. Neurochem. 1977, 28, 1175–1182. [Google Scholar] [CrossRef]
  137. Ledig, M.; Ciesielski, L.; Simler, S.; Lorentz, J.; Mandel, P. Effect of pre- and postnatal alcohol consumption on GABA levels of various brain regions in the rat offspring. Alcohol Alcohol 1988, 23, 63–67. [Google Scholar]
  138. Lee, K.; Kentroti, S.; Vernadakis, A. Differential sensitivity of cholinergic and GABAergic neurons in chick embryos treated intracerebrally with ethanol at 8 days of embryonic age. Neurochem. Res. 1992, 17, 565–569. [Google Scholar] [CrossRef]
  139. Kentroti, S.; Vernadakis, A. Effects of early in ovo administration of ethanol on expression of the GABAergic neuronal phenotype in the chick embryo. Brain Res. Dev. Brain Res. 1991, 61, 290–292. [Google Scholar] [CrossRef]
  140. Janiri, L.; Gobbi, G.; Persico, A.; Santarelli, M.; Minciacchi, D.; Tempesta, E. Alterations of neocortical neuronal responses to acetylcholine and GABA in rats born to alcohol-dependent mothers. Alcohol Alcohol 1994, 29, 611–619. [Google Scholar]
  141. Moore, D.B.; Ruygrok, A.; Walker, D.; Heaton, M. Effects of prenatal ethanol exposure on parvalbumin-expressing GABAergic neurons in the adult rat medial septum. Alcohol Clin. Exp. Res. 1997, 21, 849–856. [Google Scholar] [CrossRef]
  142. Moore, D.B.; Quintero, M.; Ruygrok, A.; Walker, D.; Heaton, M. Prenatal ethanol exposure reduces parvalbumin-immunoreactive GABAergic neuronal number in the adult rat cingulate cortex. Neurosci. Lett. 1998, 249, 25–28. [Google Scholar] [CrossRef] [PubMed]
  143. Bailey, C.D.; Brien, J.; Reynolds, J. Chronic prenatal ethanol exposure alters the proportion of GABAergic neurons in layers II/III of the adult guinea pig somatosensory cortex. Neurotoxicol. Teratol. 2004, 26, 59–63. [Google Scholar] [CrossRef] [PubMed]
  144. Miller, M.W. Effect of prenatal exposure to ethanol on glutamate and GABA immunoreactivity in macaque somatosensory and motor cortices: Critical timing of exposure. Neuroscience 2006, 138, 97–107. [Google Scholar] [CrossRef]
  145. Cuzon, V.C.; Yeh, P.; Yanagawa, Y.; Obata, K.; Yeh, H. Ethanol consumption during early pregnancy alters the disposition of tangentially migrating GABAergic interneurons in the fetal cortex. J. Neurosci. 2008, 28, 1854–1864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Skorput, A.G.; Gupta, V.; Yeh, P.; Yeh, H. Persistent Interneuronopathy in the Prefrontal Cortex of Young Adult Offspring Exposed to Ethanol In Utero. J. Neurosci. 2015, 35, 10977–10988. [Google Scholar] [CrossRef]
  147. Skorput, A.G.; Lee, S.; Yeh, P.; Yeh, H. The NKCC1 antagonist bumetanide mitigates interneuronopathy associated with ethanol exposure in utero. Elife 2019, 8, e48648. [Google Scholar] [CrossRef]
  148. Skorput, A.G.; Yeh, H.H. Chronic Gestational Exposure to Ethanol Leads to Enduring Aberrances in Cortical Form and Function in the Medial Prefrontal Cortex. Alcohol Clin. Exp. Res. 2016, 40, 1479–1488. [Google Scholar] [CrossRef] [Green Version]
  149. Larsen, Z.H.; Chander, P.; Joyner, J.; Floruta, C.; Demeter, T.; Weick, J. Effects of Ethanol on Cellular Composition and Network Excitability of Human Pluripotent Stem Cell-Derived Neurons. Alcohol Clin. Exp. Res. 2016, 40, 2339–2350. [Google Scholar] [CrossRef] [Green Version]
  150. Kenton, J.A.; Ontiveros, T.; Bird, C.; Valenzuela, C.; Brigman, J. Moderate prenatal alcohol exposure alters the number and function of GABAergic interneurons in the murine orbitofrontal cortex. Alcohol 2020, 88, 33–41. [Google Scholar] [CrossRef]
  151. Marquardt, K.; Sigdel, R.; Caldwell, K.; Brigman, J. Prenatal ethanol exposure impairs executive function in mice into adulthood. Alcohol Clin. Exp. Res. 2014, 38, 2962–2968. [Google Scholar] [CrossRef] [Green Version]
  152. Granato, A. Altered organization of cortical interneurons in rats exposed to ethanol during neonatal life. Brain Res. 2006, 1069, 23–30. [Google Scholar] [CrossRef] [PubMed]
  153. Nirgudkar, P.; Taylor, D.; Yanagawa, Y.; Valenzuela, C. Ethanol exposure during development reduces GABAergic/glycinergic neuron numbers and lobule volumes in the mouse cerebellar vermis. Neurosci. Lett. 2016, 632, 86–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Bird, C.W.; Taylor, D.; Pinkowski, N.; Chavez, G.; Valenzuela, C. Long-term Reductions in the Population of GABAergic Interneurons in the Mouse Hippocampus following Developmental Ethanol Exposure. Neuroscience 2018, 383, 60–73. [Google Scholar] [CrossRef]
  155. Coleman, L.G., Jr.; Oguz, I.; Lee, J.; Styner, M.; Crews, F. Postnatal day 7 ethanol treatment causes persistent reductions in adult mouse brain volume and cortical neurons with sex specific effects on neurogenesis. Alcohol 2012, 46, 603–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Sadrian, B.; Lopez-Guzman, M.; Wilson, D.; Saito, M. Distinct neurobehavioral dysfunction based on the timing of developmental binge-like alcohol exposure. Neuroscience 2014, 280, 204–219. [Google Scholar] [CrossRef] [PubMed]
  157. Saito, M.; Smiley, J.; Hui, M.; Masiello, K.; Betz, J.; Ilina, M.; Saito, M.; Wilson, D. Neonatal Ethanol Disturbs the Normal Maturation of Parvalbumin Interneurons Surrounded by Subsets of Perineuronal Nets in the Cerebral Cortex: Partial Reversal by Lithium. Cereb. Cortex. 2019, 29, 1383–1397. [Google Scholar] [CrossRef]
  158. Smiley, J.F.; Saito, M.; Bleiwas, C.; Masiello, K.; Ardekani, B.; Guilfoyle, D.; Gerum, S.; Wilson, D.; Vadasz, C. Selective reduction of cerebral cortex GABA neurons in a late gestation model of fetal alcohol spectrum disorder. Alcohol 2015, 49, 571–580. [Google Scholar] [CrossRef] [Green Version]
  159. Bird, C.W.; Chavez, G.; Barber, M.; Valenzuela, C. Enhancement of parvalbumin interneuron-mediated neurotransmission in the retrosplenial cortex of adolescent mice following third trimester-equivalent ethanol exposure. Sci. Rep. 2021, 11, 1716. [Google Scholar] [CrossRef]
  160. De Giorgio, A.; Comparini, S.; Intra, F.; Granato, A. Long-term alterations of striatal parvalbumin interneurons in a rat model of early exposure to alcohol. J. Neurodev. Disord. 2012, 4, 18. [Google Scholar] [CrossRef]
  161. Swartzwelder, H.S.; Farr, K.; Wilson, W.; Savage, D. Prenatal exposure to ethanol decreases physiological plasticity in the hippocampus of the adult rat. Alcohol 1988, 5, 121–124. [Google Scholar] [CrossRef]
  162. Varaschin, R.K.; Akers, K.; Rosenberg, M.; Hamilton, D.; Savage, D. Effects of the cognition-enhancing agent ABT-239 on fetal ethanol-induced deficits in dentate gyrus synaptic plasticity. J. Pharmacol. Exp. Ther. 2010, 334, 191–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. An, L.; Zhang, T. Spatial cognition and sexually dimorphic synaptic plasticity balance impairment in rats with chronic prenatal ethanol exposure. Behav. Brain Res. 2013, 256, 564–574. [Google Scholar] [CrossRef] [PubMed]
  164. Patten, A.R.; Brocardo, P.; Sakiyama, C.; Wortman, R.; Noonan, A.; Gil-Mohapel, J.; Christie, B. Impairments in hippocampal synaptic plasticity following prenatal ethanol exposure are dependent on glutathione levels. Hippocampus 2013, 23, 1463–1475. [Google Scholar] [CrossRef] [PubMed]
  165. Ito, K.; Skinkle, K.; Hicks, T. Age-dependent, steroid-specific effects of oestrogen on long-term potentiation in rat hippocampal slices. J. Physiol. 1999, 515 Pt 1, 209–220. [Google Scholar] [CrossRef]
  166. Uban, K.A.; Herting, M.; Wozniak, J.; Sowell, E.; Cifasd. Sex differences in associations between white matter microstructure and gonadal hormones in children and adolescents with prenatal alcohol exposure. Psychoneuroendocrinology 2017, 83, 111–121. [Google Scholar] [CrossRef]
  167. Sliwowska, J.H.; Comeau, W.; Bodnar, T.; Ellis, L.; Weinberg, J. Prenatal Alcohol Exposure and Pair Feeding Differentially Impact Puberty and Reproductive Development in Female Rats: Role of the Kisspeptin System. Alcohol Clin. Exp. Res. 2016, 40, 2368–2376. [Google Scholar] [CrossRef] [Green Version]
  168. Richardson, D.P.; Byrnes, M.; Brien, J.; Reynolds, J.; Dringenberg, H. Impaired acquisition in the water maze and hippocampal long-term potentiation after chronic prenatal ethanol exposure in the guinea-pig. Eur. J. Neurosci. 2002, 16, 1593–1598. [Google Scholar] [CrossRef]
  169. Izumi, Y.; Kitabayashi, R.; Funatsu, M.; Izumi, M.; Yuede, C.; Hartman, R.; Wozniak, D.; Zorumski, C. A single day of ethanol exposure during development has persistent effects on bi-directional plasticity, N-methyl-D-aspartate receptor function and ethanol sensitivity. Neuroscience 2005, 136, 269–279. [Google Scholar] [CrossRef]
  170. Puglia, M.P.; Valenzuela, C.F. Repeated third trimester-equivalent ethanol exposure inhibits long-term potentiation in the hippocampal CA1 region of neonatal rats. Alcohol 2010, 44, 283–290. [Google Scholar] [CrossRef] [Green Version]
  171. Subbanna, S.; Basavarajappa, B.S. Pre-administration of G9a/GLP inhibitor during Synaptogenesis Prevents Postnatal Ethanol-induced LTP Deficits and Neurobehavioral Abnormalities in Adult Mice. Exp. Neurol. 2014, 261, 34–43. [Google Scholar] [CrossRef] [Green Version]
  172. Subbanna, S.; Basavarajappa, B.S. Binge-like Prenatal Ethanol Exposure Causes Impaired Cellular Differentiation in the Embryonic Forebrain and Synaptic and Behavioral Defects in Adult Mice. Brain Sci. 2022, 12, 793. [Google Scholar] [CrossRef] [PubMed]
  173. De La Fuente-Ortega, E.; Plaza-Briceno, W.; Vargas-Robert, S.; Haeger, P. Prenatal Ethanol Exposure Misregulates Genes Involved in Iron Homeostasis Promoting a Maladaptation of Iron Dependent Hippocampal Synaptic Transmission and Plasticity. Front. Pharmacol. 2019, 10, 1312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Helfer, J.L.; White, E.; Christie, B. Enhanced deficits in long-term potentiation in the adult dentate gyrus with 2nd trimester ethanol consumption. PLoS ONE 2012, 7, e51344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Patten, A.R.; Brocardo, P.; Christie, B. Omega-3 supplementation can restore glutathione levels and prevent oxidative damage caused by prenatal ethanol exposure. J. Nutr. Biochem. 2013, 24, 760–769. [Google Scholar] [CrossRef]
  176. Varaschin, R.K.; Rosenberg, M.; Hamilton, D.; Savage, D. Differential effects of the histamine H(3) receptor agonist methimepip on dentate granule cell excitability, paired-pulse plasticity and long-term potentiation in prenatal alcohol-exposed rats. Alcohol Clin. Exp. Res. 2014, 38, 1902–1911. [Google Scholar] [CrossRef]
  177. Silvestre de Ferron, B.; Vilpoux, C.; Kervern, M.; Robert, A.; Antol, J.; Naassila, M.; Pierrefiche, O. Increase of KCC2 in hippocampal synaptic plasticity disturbances after perinatal ethanol exposure. Addict. Biol. 2017, 22, 1870–1882. [Google Scholar] [CrossRef]
  178. Fontaine, C.J.; Pinar, C.; Yang, W.; Pang, A.; Suesser, K.; Choi, J.; Christie, B. Impaired Bidirectional Synaptic Plasticity in Juvenile Offspring Following Prenatal Ethanol Exposure. Alcohol Clin. Exp. Res. 2019, 43, 2153–2166. [Google Scholar] [CrossRef]
  179. Grafe, E.L.; Wade, M.; Hodson, C.; Thomas, J.; Christie, B. Postnatal Choline Supplementation Rescues Deficits in Synaptic Plasticity Following Prenatal Ethanol Exposure. Nutrients 2022, 14, 2004. [Google Scholar] [CrossRef]
  180. Titterness, A.K.; Christie, B.R. Long-term depression in vivo: Effects of sex, stress, diet, and prenatal ethanol exposure. Hippocampus 2008, 18, 481–491. [Google Scholar] [CrossRef] [PubMed]
  181. Lee, F.J.; Wang, Y.; Liu, F. Direct receptor cross-talk can mediate the modulation of excitatory and inhibitory neurotransmission by dopamine. J. Mol. Neurosci. 2005, 26, 245–252. [Google Scholar] [CrossRef] [PubMed]
  182. Lucchi, L.; Covelli, V.; Petkov, V.; Spano, P.; Trabucchi, M. Effects of ethanol, given during pregnancy, on the offspring dopaminergic system. Pharmacol. Biochem. Behav. 1983, 19, 567–570. [Google Scholar] [CrossRef] [PubMed]
  183. Lucchi, L.; Covelli, V.; Spano, P.; Trabucchi, M. Acute ethanol administration during pregnancy: Effects on central dopaminergic transmission in rat offspring. Neurobehav. Toxicol. Teratol. 1984, 6, 19–21. [Google Scholar] [PubMed]
  184. Rathbun, W.; Druse, M.J. Dopamine, serotonin, and acid metabolites in brain regions from the developing offspring of ethanol-treated rats. J. Neurochem. 1985, 44, 57–62. [Google Scholar] [CrossRef] [PubMed]
  185. Druse, M.J.; Tajuddin, N.; Kuo, A.; Connerty, M. Effects of in utero ethanol exposure on the developing dopaminergic system in rats. J. Neurosci. Res. 1990, 27, 233–240. [Google Scholar] [CrossRef] [PubMed]
  186. Boggan, W.O.; Xu, W.; Shepherd, C.; Middaugh, L. Effects of prenatal ethanol exposure on dopamine systems in C57BL/6J mice. Neurotoxicol. Teratol. 1996, 18, 41–48. [Google Scholar] [CrossRef] [PubMed]
  187. Nowak, P.; Dabrowska, J.; Bortel, A.; Izabela, B.; Kostrzewa, R.; Brus, R. Prenatal cadmium and ethanol increase amphetamine-evoked dopamine release in rat striatum. Neurotoxicol. Teratol. 2006, 28, 563–572. [Google Scholar] [CrossRef]
  188. Tattoli, M.; Cagiano, R.; Gaetani, S.; Ghiglieri, V.; Giustino, A.; Mereu, G.; Trabace, L.; Cuomo, V. Neurofunctional effects of developmental alcohol exposure in alcohol-preferring and alcohol-nonpreferring rats. Neuropsychopharmacology 2001, 24, 691–705. [Google Scholar] [CrossRef]
  189. Shen, R.Y.; Hannigan, J.; Chiodo, L. The effects of chronic amphetamine treatment on prenatal ethanol-induced changes in dopamine receptor function: Electrophysiological findings. J. Pharmacol. Exp. Ther. 1995, 274, 1054–1060. [Google Scholar]
  190. Shen, R.Y.; Chiodo, L.A. The effects of in utero ethanol administration on the electrophysiological activity of rat nigrostriatal dopaminergic neurons. Brain Res. 1993, 624, 216–222. [Google Scholar] [CrossRef]
  191. Xu, C.; Shen, R.Y. Amphetamine normalizes the electrical activity of dopamine neurons in the ventral tegmental area following prenatal ethanol exposure. J. Pharmacol. Exp. Ther. 2001, 297, 746–752. [Google Scholar]
  192. Choong, K.; Shen, R. Prenatal ethanol exposure alters the postnatal development of the spontaneous electrical activity of dopamine neurons in the ventral tegmental area. Neuroscience 2004, 126, 1083–1091. [Google Scholar] [CrossRef] [PubMed]
  193. Wang, J.; Haj-Dahmane, S.; Shen, R. Effects of prenatal ethanol exposure on the excitability of ventral tegmental area dopamine neurons in vitro. J. Pharmacol. Exp. Ther. 2006, 319, 857–863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Zhou, R.; Wang, S.; Zhu, X. Prenatal ethanol exposure alters synaptic plasticity in the dorsolateral striatum of rat offspring via changing the reactivity of dopamine receptor. PLoS One 2012, 7, e42443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Kroner, S.; Rosenkranz, J.; Grace, A.; Barrionuevo, G. Dopamine modulates excitability of basolateral amygdala neurons in vitro. J. Neurophysiol. 2005, 93, 1598–1610. [Google Scholar] [CrossRef]
  196. Diaz, M.R.; Chappell, A.; Christian, D.; Anderson, N.; McCool, B. Dopamine D3-like receptors modulate anxiety-like behavior and regulate GABAergic transmission in the rat lateral/basolateral amygdala. Neuropsychopharmacology 2011, 36, 1090–1103. [Google Scholar] [CrossRef]
  197. Diaz, M.R.; Jotty, K.; Locke, J.; Jones, S.; Valenzuela, C. Moderate Alcohol Exposure during the Rat Equivalent to the Third Trimester of Human Pregnancy Alters Regulation of GABAA Receptor-Mediated Synaptic Transmission by Dopamine in the Basolateral Amygdala. Front. Pediatr. 2014, 2, 46. [Google Scholar] [CrossRef] [Green Version]
  198. Hausknecht, K.; Shen, Y.; Wang, R.; Haj-Dahmane, S.; Shen, R. Prenatal Ethanol Exposure Persistently Alters Endocannabinoid Signaling and Endocannabinoid-Mediated Excitatory Synaptic Plasticity in Ventral Tegmental Area Dopamine Neurons. J. Neurosci. 2017, 37, 5798–5808. [Google Scholar] [CrossRef]
  199. Cuzon Carlson, V.C.; Gremel, C.; Lovinger, D. Gestational alcohol exposure disrupts cognitive function and striatal circuits in adult offspring. Nat. Commun. 2020, 11, 2555. [Google Scholar] [CrossRef]
  200. Oubraim, S.; Wang, R.; Hausknecht, K.; Kaczocha, M.; Shen, R.; Haj-Dahmane, S. Prenatal ethanol exposure causes anxiety-like phenotype and alters synaptic nitric oxide and endocannabinoid signaling in dorsal raphe nucleus of adult male rats. Transl. Psychiatry 2022, 12, 440. [Google Scholar] [CrossRef]
  201. Druse, M.J.; Kuo, A.; Tajuddin, N. Effects of in utero ethanol exposure on the developing serotonergic system. Alcohol Clin. Exp. Res. 1991, 15, 678–684. [Google Scholar] [CrossRef]
  202. Sari, Y.; Powrozek, T.; Zhou, F. Alcohol deters the outgrowth of serotonergic neurons at midgestation. J. Biomed. Sci. 2001, 8, 119–125. [Google Scholar] [CrossRef] [PubMed]
  203. Zhou, F.C.; Sari, Y.; Powrozek, T. Fetal alcohol exposure reduces serotonin innervation and compromises development of the forebrain along the serotonergic pathway. Alcohol Clin. Exp. Res. 2005, 29, 141–149. [Google Scholar] [CrossRef] [PubMed]
  204. Zhou, F.C.; Sari, Y.; Zhang, J.; Goodlett, C.; Li, T. Prenatal alcohol exposure retards the migration and development of serotonin neurons in fetal C57BL mice. Brain Res. Dev. Brain Res. 2001, 126, 147–155. [Google Scholar] [CrossRef] [PubMed]
  205. Kim, E.K.; Lee, M.; Kim, H.; Sim, Y.; Shin, M.; Lee, S.; Yang, H.; Chang, H.; Lee, T.; Jang, M.; et al. Maternal ethanol administration inhibits 5-hydroxytryptamine synthesis and tryptophan hydroxylase expression in the dorsal raphe of rat offspring. Brain Dev. 2005, 27, 472–476. [Google Scholar] [CrossRef] [PubMed]
  206. Madarnas, C.; Villalba, N.; Soriano, D.; Brusco, A. Anxious Behavior of Adult CD1 Mice Perinatally Exposed to Low Concentrations of Ethanol Correlates With Morphological Changes in Cingulate Cortex and Amygdala. Front. Behav. Neurosci. 2020, 14, 92. [Google Scholar] [CrossRef]
  207. Marsicano, G.; Lutz, B. Expression of the cannabinoid receptor CB1 in distinct neuronal subpopulations in the adult mouse forebrain. Eur. J. Neurosci. 1999, 11, 4213–4225. [Google Scholar] [CrossRef]
  208. Soria-Gomez, E.; Bellocchio, L.; Reguero, L.; Lepousez, G.; Martin, C.; Bendahmane, M.; Ruehle, S.; Remmers, F.; Desprez, T.; Matias, I.; et al. The endocannabinoid system controls food intake via olfactory processes. Nat. Neurosci. 2014, 17, 407–415. [Google Scholar] [CrossRef]
  209. Ohno-Shosaku, T.; Maejima, T.; Kano, M. Endogenous cannabinoids mediate retrograde signals from depolarized postsynaptic neurons to presynaptic terminal. Neuron 2001, 29, 729–738. [Google Scholar] [CrossRef]
  210. Wilson, R.I.; Nicoll, R.A. Endogenous cannabinoids mediate retrograde signalling at hippocampal synapses. Nature 2001, 410, 588–592. [Google Scholar] [CrossRef]
  211. Basavarajappa, B.S.; Joshi, V.; Shivakumar, M.; Subbanna, S. Distinct Functions of Endogenous Cannabinoid System in Alcohol Abuse Disorders. Br. J. Pharmacol. 2019, 176, 3085–3109. [Google Scholar] [CrossRef]
  212. Hoffman, A.F.; Hwang, E.; Lupica, C. Impairment of Synaptic Plasticity by Cannabis, Delta(9)-THC, and Synthetic Cannabinoids. Cold Spring Harb. Perspect. Med. 2021, 11, a039743. [Google Scholar] [CrossRef]
  213. Fernandez-Ruiz, J.; Berrendero, F.; Hernandez, M.; Ramos, J. The endogenous cannabinoid system and brain development. Trends Neurosci. 2000, 23, 14–20. [Google Scholar] [CrossRef] [PubMed]
  214. de Salas-Quiroga, A.; Diaz-Alonso, J.; Garcia-Rincon, D.; Remmers, F.; Vega, D.; Gomez-Canas, M.; Lutz, B.; Guzman, M.; Galve-Roperh, I. Prenatal exposure to cannabinoids evokes long-lasting functional alterations by targeting CB1 receptors on developing cortical neurons. Proc. Natl. Acad. Sci. USA 2015, 112, 13693–13698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Stefanis, N.C.; Delespaul, P.; Henquet, C.; Bakoula, C.; Stefanis, C.; Van Os, J. Early adolescent cannabis exposure and positive and negative dimensions of psychosis. Addiction 2004, 99, 1333–1341. [Google Scholar] [CrossRef] [PubMed]
  216. Wu, C.S.; Jew, C.; Lu, H. Lasting impacts of prenatal cannabis exposure and the role of endogenous cannabinoids in the developing brain. Future Neurol. 2011, 6, 459–480. [Google Scholar] [CrossRef] [Green Version]
  217. Hansen, H.H.; Krutz, B.; Sifringer, M.; Stefovska, V.; Bittigau, P.; Pragst, F.; Marsicano, G.; Lutz, B.; Ikonomidou, C. Cannabinoids enhance susceptibility of immature brain to ethanol neurotoxicity. Ann. Neurol. 2008, 64, 42–52. [Google Scholar] [CrossRef]
  218. Noel, M.; Norris, E.; Strickland, S. Tissue plasminogen activator is required for the development of fetal alcohol syndrome in mice. Proc. Natl. Acad. Sci. USA 2011, 108, 5069–5074. [Google Scholar] [CrossRef]
  219. Sadrian, B.; Subbanna, S.; Wilson, D.; Basavarajappa, B.; Saito, M. Lithium prevents long-term neural and behavioral pathology induced by early alcohol exposure. Neuroscience 2012, 206, 122–135. [Google Scholar] [CrossRef] [Green Version]
  220. Wilson, D.A.; Peterson, J.; Basavaraj, B.; Saito, M. Local and regional network function in behaviorally relevant cortical circuits of adult mice following postnatal alcohol exposure. Alcohol. Clin. Exp. Res. 2011, 35, 1974–1984. [Google Scholar] [CrossRef] [Green Version]
  221. Boa-Amponsem, O.; Zhang, C.; Mukhopadhyay, S.; Ardrey, I.; Cole, G. Ethanol and cannabinoids interact to alter behavior in a zebrafish fetal alcohol spectrum disorder model. Birth Defects Res. 2019, 111, 775–788. [Google Scholar] [CrossRef] [PubMed]
  222. Fish, E.W.; Murdaugh, L.; Zhang, C.; Boschen, K.; Boa-Amponsem, O.; Mendoza-Romero, H.; Tarpley, M.; Chdid, L.; Mukhopadhyay, S.; Cole, G.; et al. Cannabinoids Exacerbate Alcohol Teratogenesis by a CB1-Hedgehog Interaction. Sci. Rep. 2019, 9, 16057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. McCorvy, J.D.; Roth, B.L. Structure and function of serotonin G protein-coupled receptors. Pharmacol. Ther. 2015, 150, 129–142. [Google Scholar] [CrossRef] [Green Version]
  224. Ugedo, L.; De Deurwaerdere, P. Serotonergic control of the glutamatergic neurons of the subthalamic nucleus. Prog. Brain Res. 2021, 261, 423–462. [Google Scholar] [PubMed]
  225. Rahaman, S.M.; Chowdhury, S.; Mukai, Y.; Ono, D.; Yamaguchi, H.; Yamanaka, A. Functional Interaction Between GABAergic Neurons in the Ventral Tegmental Area and Serotonergic Neurons in the Dorsal Raphe Nucleus. Front. Neurosci. 2022, 16, 877054. [Google Scholar] [CrossRef] [PubMed]
  226. Frederick, A.L.; Stanwood, G.D. Drugs, biogenic amine targets and the developing brain. Dev. Neurosci. 2009, 31, 7–22. [Google Scholar] [CrossRef]
Figure 1. The diagram illustrates the structural diversity of AMPA and NMDA receptor subunits. Multiple subunits have been characterized in these receptor classes that bind either glutamate or glycine/D-serine. Subunits of AMPA receptors can form functional homomers and heteromers, but GluA2-containing heteromers are the most common in the brain [86]. AMPA receptors are classified as GluA2-containing (Ca2+-impermeable and insensitive to polyamines) and GluA2-lacking (Ca2+-permeable and blocked by polyamines). NMDA receptors exist as heteromeric receptors with two GluN1 subunits and two GluN2 subunits (GluN1/2) or two GluN3 subunits (GluN1/3). NMDA receptor activation requires the simultaneous binding of glycine (or D-serine) and glutamate to relieve the voltage-dependent Mg2+ block and allow the flow of inward Ca2+ [86].
Figure 1. The diagram illustrates the structural diversity of AMPA and NMDA receptor subunits. Multiple subunits have been characterized in these receptor classes that bind either glutamate or glycine/D-serine. Subunits of AMPA receptors can form functional homomers and heteromers, but GluA2-containing heteromers are the most common in the brain [86]. AMPA receptors are classified as GluA2-containing (Ca2+-impermeable and insensitive to polyamines) and GluA2-lacking (Ca2+-permeable and blocked by polyamines). NMDA receptors exist as heteromeric receptors with two GluN1 subunits and two GluN2 subunits (GluN1/2) or two GluN3 subunits (GluN1/3). NMDA receptor activation requires the simultaneous binding of glycine (or D-serine) and glutamate to relieve the voltage-dependent Mg2+ block and allow the flow of inward Ca2+ [86].
Cells 12 00442 g001
Figure 2. The outline illustrates the structural diversity of GABAA receptor subunits. GABAA ionotropic receptors have been characterized as heteromers consisting of five subunits, most generally two α’s, two β’s, and one γ (α2β2γ). Each subunit consists of many subtypes (α1-6, β1-3, and γ1-3) and exhibits different properties, distributions in the brain, and a wide range of activities in response to pharmacological agents.
Figure 2. The outline illustrates the structural diversity of GABAA receptor subunits. GABAA ionotropic receptors have been characterized as heteromers consisting of five subunits, most generally two α’s, two β’s, and one γ (α2β2γ). Each subunit consists of many subtypes (α1-6, β1-3, and γ1-3) and exhibits different properties, distributions in the brain, and a wide range of activities in response to pharmacological agents.
Cells 12 00442 g002
Figure 3. Alterations in glutamatergic and GABAergic neurons and their modulators were found in FASD models. Developmental alcohol exposure causes persistent changes in glutamatergic and GABAergic neuron function, causing synaptic plasticity defects that result in many psychiatric disorders found in FASD, including cognitive and social interaction behavioral impairments.
Figure 3. Alterations in glutamatergic and GABAergic neurons and their modulators were found in FASD models. Developmental alcohol exposure causes persistent changes in glutamatergic and GABAergic neuron function, causing synaptic plasticity defects that result in many psychiatric disorders found in FASD, including cognitive and social interaction behavioral impairments.
Cells 12 00442 g003
Table 2. Summary of developmental alcohol exposure effects on GABAergic neurotransmitter system.
Table 2. Summary of developmental alcohol exposure effects on GABAergic neurotransmitter system.
Alcohol ModelBACTissue/RegionEffects
GD1–14 (Alc liquid diet) (A Rats)NDWBIncreased GABA (GD18&21) [76,136].
GD1–PD14 (10% Alc sol, Oral (WS Rats)360 mg/dL (PD21)THA, Pons, CBL, FC, OLB, ACIncreased GABA (FC, OLB, AC&Amy); Reduced GABA (THA, Pons, CBL&HP); (PD21) [137].
E1–E3 (10 mg/50 µL) (Chick embryo)12 mMWBIncreased GAD activity (E8) [138,139].
GD15–18 (2.4 g/kg/4 times/day) (SD rats)NDCortical neuronsIncreased GABAergic responses (PD90) [140].
GD1–20 (5% Alc liquid diet) (SD Rats)83 mg/dL, maternalmFC&HPIncreased GABAA receptor-stimulated Cl-flux [131].
GD0–21 (Alc liquid diet)
(LE Rats)
161 mg/dL, maternalMS&ACCReduced GABAergic (PV+) IN (PD60) [141,142].
GD2–67 (4 g/kg, oral) (Pigs)327 mg/dL, maternalSCReduced GAD+ cells (PD220) [143]
GD3–42 &3-168 (1.8 g/kg/day, intragastric) (Pigs)234 mg/dL, maternalSC and MCReduced GABA+ ve neurons [144].
GD1–14.5 (Alc liquid diet) (C57 mice)25 mg/dL, maternalCorticle anlagePremature GABAergic interneuron tangential migration; Increased GABA and GABA sensitivity [145].
GD13.5–16.5 (5% Alc liquid diet) (C57mice)80 mg/dL, maternalEmbryo and adult mPFCIncreased median ganglionic eminence-derived IN (E16.5) [146]; Increased PV+ IN (mPFC) & potentiated GABAAR transmission (PN) [146].
GD13.5–16.5 (5% Alc liquid diet) (Nkx2.1-Cre mice)80 mg/dL, maternalPFCIncreased depolarizing action of GABAAR in migrating neurons (PD58) [146,147].
GD1–20 (4 g/kg/day) (C57 mice)80 mg/dL, maternalOFCIncreased spontaneous IPSCs amplitude and area in PYNs (PD60) [150].
PD2–PD6 (Alc vapor) (WS Rats)206 mg/dL, maternalMC&SCIncreased CR+ ve INs;
Reduced CB+ ve INs (PD60) [152];
Reduced PV+ ve Ins (ST) [152].
GD12–19 & PD2–9 (Alc vapor) (Venus-VGAT-mouse line)330 mg/dL, PD7–8CBLReduced INs (PD16) [153].
PD2–9 (Alc vapor) (Venus-VGAT-mouse)221 mg/dL, PD9HPReduced INs (PD90) [154].
PD7 (Alc vapor) (Venus-VGAT-mouse line)297 mg/dL, PD7HPReduced INs (PD90) [154].
PD7 (2.5 g Alc × 2times) (C57 Mice))500 mg/dL, PD7mPFC, HP&PCReduced PV+ ve INs (PD90) [155,156,157].
PD7 (2.5 g Alc × 2times) (C57 Mice))500 mg/dL, PD7NCReduced PV+ ve & CA+ ve INs (PD90) [158].
PD7 (Alc vapor) (Venus-VGAT-mouse line)297 mg/dL, PD7RSCEnhancement of PV+ ve INs-mediated
Neurotransmission (PD40–60) [159].
ND, not determined.
Table 3. Summary of developmental alcohol exposure effects on hippocampal LTP and LTD in offspring.
Table 3. Summary of developmental alcohol exposure effects on hippocampal LTP and LTD in offspring.
Alcohol ModelBACHP SlicesEffects
GD1–22 (Alc liquid diet) (Rats)31 mg/dL, maternalCA1Reduced LTP (PD50–70) [161].
GD2–67 (Oral Alc, 4 g/kg/day) (Pigs)416 mg/dL, maternalCA1Reduced LTP (PD40–50) [168].
PD7 (2.5 g/kg × 2/day) (C57 mice)500 mg/dL, PD7CA1Reduced LTP (PD30–32) [169].
PD2–9 (Alc vapor) (SD Rats)395 mg/dL, PD9CA1Reduced LTP (PD7–9) [170]
GD1–22 (Alc liquid diet) (LE Rats)87 mg/dL, maternalCA1Reduced LTP in males; Increased LTP in females (PD30–35) [162].
GD0–22 (Alc 4 g/kg/day) (W Rats)NDCA1Reduced LTP in males and increased in females (PD36) [163].
GD1–22 (Alc liquid diet) (SD Rats)135 mg/dL, maternalCA1Reduced LTP in males and no change in females (PD55–65) [164].
PD7 (2.5 g/kg/twice) (C57 mice)490 mg/dL, PD7CA1Reduced LTP (PD90) [43,45,74,75,171].
GD8&12 (1.75 g/kg/twice) (C57 mice)300 mg/dL, maternalCA1Reduced LTP (PD90) [172].
GD5–PD7 (10% Alc sol) (SD rats)NDCA1Reduced LTP (PD17–30) [173].
PD7 (2.5 g/kg × 2/day) (C57 mice)500 mg/dL, PD7CA1Abolished LTD (PD30–32) [169].
GD1–22&lactation (10% alcohol sol) (SD rats)NDCA1Increased LTD (PD50–52) [111].
GD0–22 (Alc 4 g/kg/day) (WS Rats)NDCA1Reduced LTD in males and increased in females (PD36) [163].
GD8&12 (1.75 g/kg/twice) (C57 mice)300 mg/dL, maternalCA1Reduced LTD in males (PD90) [172].
GD1–22 (Alc liquid diet) (LE Rats)84 mg/dL, maternalDGReduced LTP (PD105–140) [162].
GD11–21 (Alc liquid diet)142 mg/dLDGReduced LTP in males(PD50–70) [174].
GD1–22 (Alc liquid diet) (SD Rats)135 mg/dL, maternalDGReduced LTP in males and no change in females (PD55–70) [175].
GD1–22 (Alc 5%liquid diet) (SD Rats)146 mg/dL, maternalDGReduced LTP in males and no change in females (PD55–70) [108].
GD1–22 (5% Alc in water) (LE Rats)84 mg/dL, maternalDGReduced LTP (PD105–140) [176].
GD1–PD14 (10% Alc sol.) (SD Rats)100 mg/dL, PD7DGReduced LTP (PD45–55) [177].
GD1–22 (Alc liquid diet) (SD Rats)80–180 mg/dL, maternalDGReduced LTP in males and females (PD21–28) [178].
GD1–PD7 (10% Alc sol) (SD Rats)62 mg/dL, maternalCA1 & DGReduced LTP (PD21–60) [126].
GD1–22 (Alc liquid diet) (SD Rats)80–180 mg/dL, maternalDGReduced LTP in males only (PD31–35) [179].
GD1–PD14 (10% Alc sol.)(SD Rats)100 mg/dL, PD7DGFacilitated LTD (PD45–55) [177].
GD1–22 (Alc liquid diet) (SD Rats)80–180 mg/dL, maternalDGReduced LTD in females only (PD21–28) [178].
Table 4. Summary of the influence of developmental alcohol on modulators of glutamatergic and GABAergic neurotransmitter system.
Table 4. Summary of the influence of developmental alcohol on modulators of glutamatergic and GABAergic neurotransmitter system.
Alcohol ModelBACTissue/RegionEffects
GD6–20 (alc liquid diet) (LE Rats)NDVTASupersensitive DA autoreceptors (PD90–120) [189].
GD6–20 (alc liquid diet) (LE Rats)NDNSDAReduced DA receptor functions (P90–120) [190].
GD8–20 (3 g/kg/twice/day) (SD rats)281–341 mg/dL, maternalVTAReduced the number of spontaneously active DA neurons (P90) [191].
GD8–20 (3 g/kg/twice a day) (SD rats)281–341 mg/dL, maternalVTA-DAEnhanced GluA3 (PD14–84) [79];
Enhanced EPSCs strength (PD14–84) [79].
GD7–20 (6g/kg/day) (SD Rats)302–331 mg/dL, maternalDL-STEnhanced D1R function (PD30) [194].
PD2–12 (Alc vapor, 4 h/day) (SD Rats)23 mg/dL, pupsBLADecreased DA; Reduced D1R-mediated potentiation of sIPSCs; Impaired D3R-mediated suppression of mIPSCs
(PD40–50) [197].
PD7 (2.5g/kg/twice) (C57 mice)490 mg/dL, PD7HP&NCEnhanced AEA levels, NAPE-PLD, GDE & CB1 expression (PD7) [45].
GD8–20 (3 g/kg/twice a day) (SD rats)281–341 mg/dL, maternalVTA-DAImpaired eCB-LTD (PD28–70) [198].
GD0–20 (Alc vapor)
(C57 mice)
84 mg/dL, maternalDLS-MSNsIncreased excitability of MS neurons; Increased eCB tone (PD90) [199].
GD8–20 & PD0–10 (3 g/kg/twice a day) (SD rats)281–341 mg/dL, maternalVTA-DAReduced CB1 mRNA expression (PD60–70) [200].
GD1–20 (Alc liquid diet)
(SD Rats)
NDMC & SSCReduced 5-HT, 5-HIAA &5-HT1 receptors (PD19) [184,201].
GD8–15 (Alc liquid diet) (C57 mice)NDMFB, MR, DR, HP, etc.Reduced 5-HT (E15/18) [202,203,204].
GD15–20 (0.5–2g/kg/day) (SD Rats)3.32–106 mg/dL, maternalDRReduced 5-HT and TPH (PD37) [205].
GD1–20-PD20 (6% Alc) (CD1 mice)73–102 mg/dL (PD21)Amy&CCReduced 5-HT (Amy&CC); Reduced 5-HT1R (CC); Increased 5-HTR (Amy) (PD77–84) [206].
GD8–20 & PD0–10 (3 g/kg/twice a day) (SD rats)281–341 mg/dL, maternalVTA-DAEnhanced the electrical activity of DRn 5-HT neurons (PD60–70) [200].
ND, not determined.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Basavarajappa, B.S.; Subbanna, S. Synaptic Plasticity Abnormalities in Fetal Alcohol Spectrum Disorders. Cells 2023, 12, 442. https://doi.org/10.3390/cells12030442

AMA Style

Basavarajappa BS, Subbanna S. Synaptic Plasticity Abnormalities in Fetal Alcohol Spectrum Disorders. Cells. 2023; 12(3):442. https://doi.org/10.3390/cells12030442

Chicago/Turabian Style

Basavarajappa, Balapal S., and Shivakumar Subbanna. 2023. "Synaptic Plasticity Abnormalities in Fetal Alcohol Spectrum Disorders" Cells 12, no. 3: 442. https://doi.org/10.3390/cells12030442

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop