Abstract
There is a considerable body of literature demonstrating that adolescence is a unique age period, which includes rapid and dramatic maturation of behavioral, cognitive, hormonal and neurobiological systems. Most notably, adolescence is also a period of unique responsiveness to alcohol effects, with both hyposensitivity and hypersensitivity observed to the various effects of alcohol. Multiple neurotransmitter systems are undergoing fine-tuning during this critical period of brain development, including those that contribute to the rewarding effects of drugs of abuse. The role of developmental maturation of the γ-amino-butyric acid (GABA) system, however, has received less attention in contributing to age-specific alcohol sensitivities. This review integrates GABA findings from human magnetic resonance spectroscopy studies as they may translate to understanding adolescent-specific responsiveness to alcohol effects. Better understanding of the vulnerability of the GABA system both during adolescent development, and in psychiatric conditions that include alcohol dependence, could point to a putative mechanism, boosting brain GABA, that may have increased effectiveness for treating alcohol abuse disorders.
Keywords: adolescent, alcohol sensitivity, brain, MRS, GABA
1. Introduction
Alcohol use is typically initiated during the period of adolescence in humans (Bates & Labouvie, 1997; Hingson, Heeren, & Winter, 2006a), a critical age span that overlaps with significant brain maturational processes the adolescent brain more vulnerable to alcohol exposure than their more “neurobiologically adult” counterparts (Bava & Tapert, 2010; Blakemore, 2012; Clark, Thatcher, & Tapert, 2008). There are several behavioral features associated with adolescence, including but not limited to increased impulsiveness and risk taking, and suboptimal decision-making skills, which could contribute to increased vulnerabilities for developing heavy alcohol and drug use (Nixon & McClain, 2010; Spear, 2011a; Tapert & Schweinsburg, 2005) or early manifestation of psychiatric illnesses (Gogtay & Thompson, 2010; Paus, Keshavan, & Giedd, 2008) during this age span. These behavioral features have neurobiological underpinnings that include structural, functional and neurochemical developmental brain changes that are unique to adolescence, that are particularly prominent in the frontal lobe and that are commensurate with developmental improvements in executive functioning skills (Casey, Galvan, & Hare, 2005; Gogtay, et al., 2004). These features observed in adolescents overlap with similar features observed in cohorts of adult patients with alcohol abuse disorders and psychiatric illnesses, namely risk-taking, impulsiveness, and differences in brain tissue volume, functional brain activation and neurochemistry. This overlap collectively highlights the vulnerability of the adolescent brain to the early manifestation of addictive disorders and psychiatric conditions. Accordingly, characterizing the ontogeny of multiple neurotransmitter systems has been the subject of numerous investigations (for review, see (Spear, 2000), however more recently, investigation into the development of the inhibitory neural system, γ-amino-butyric acid (GABA), is gaining significant attention as a potential moderator of human developmental changes in impulse control, self-regulation and decision-making (Silveri, et al., 2013). Parallel to studies investigating contributions of brain GABA to functioning during adolescence is an accumulation of GABA findings based on studies conducted in healthy adults (Boy, et al., 2010; Goto, et al., 2010; Northoff, et al., 2007; Stagg, 2013; Sumner, Edden, Bompas, Evans, & Singh, 2010) and in adult patients with psychiatric illnesses (Bhagwagar, et al., 2008; Ham, et al., 2007; Long, et al., 2013; Plante, Jensen, Schoerning, & Winkelman, 2012a; Pollack, Jensen, Simon, Kaufman, & Renshaw, 2008; Rosso, et al., 2013; Simpson, et al., 2012; Streeter, et al., 2005; Yoon, et al., 2010). Notably, the role of GABA in alcohol responsiveness has been well established in animal models, and GABA-related alterations have been investigated in humans with alcohol abuse disorders (Abe, et al., 2012; Behar, et al., 1999; Mason, et al., 2006; Mon, Durazzo, & Meyerhoff, 2012; Silveri, et al., 2014). Together these studies implicate in vivo brain GABA alterations in disease symptomatologies, particularly in alcohol dependence, alcoholism vulnerability, and pharmacotherapies available for treating alcoholism (Krystal, et al., 2006) (Figure 1).
Figure 1. Clinical evidence of low in vivo brain GABA levels, measured using MRS in adolescents, alcohol abuse disorders and psychiatric conditions.
Schematic highlighting the overlap of low brain GABA levels measured using MRS in adolescents, adults with alcohol abuse disorders and adults with psychiatric disorders (e.g., major depressive disorder, anxiety disorders).
Examples are provided that span domains of motor control, anxiety, and cognition, which permit an opportunity to draw parallels between preclinical studies of adolescent-specific responsiveness to alcohol and clinical studies of brain GABA levels, with a goal of inspiring novel lines of research investigations in alcohol abuse research.
2. Setting the Stage
2.1. Epidemiology of Adolescent Alcohol Use
Excessive alcohol consumption is the third leading cause of preventable death in the United States (Mokdad, Marks, Stroup, & Gerberding, 2004). The onset of alcohol use frequently occurs during adolescence, increasing in prevalence of use from 3% at age 12 to 68% at age 21 (SAMHSA, 2010a). As prevalence of use increases with age, drinking patterns and quantity of alcohol consumed likewise increase. A recent national survey indicated that 508,329 adolescents aged 12–17 years drank alcohol on at least one day during the past year, averaging 4.6 drinks consumed per day of drinking (SAMHSA, 2010b). Thus, patterns of alcohol drinking often reach heavy episodic, or binge-like levels during adolescence. Binge drinking is generally defined as a pattern of drinking that raises blood alcohols levels to 0.08 gram% or greater over a two hour period, which typically translates to 5+ drinks for adult men and 4+ drinks for adult women (NIAAA, 2004). In adolescents, 7% of 8th graders, 16% of 10th graders, and 23% of 12th graders report having drank 5+ drinks in a row within the two weeks prior to survey completion (Johnston, O’Malley, Bachman, & Schulenberg, 2011). Furthermore, prevalence of alcohol dependence in adulthood is strongly influenced by age of first alcohol use: there is a 38.0% chance of developing alcohol dependence later in life if drinking onset occurs before age 15. Importantly, the rate of prevalence (likelihood) subsequently decreases with each year that the onset of alcohol consumption is delayed during adolescence, falling to 10% (but not zero) if individuals wait until the legal age to drink in the United States (21 years old) (Grant & Dawson, 1997; Hingson, Heeren, & Winter, 2006b).
2.2. Adolescent Brain Development: Focus on GABA
The age of onset of alcohol use, the rapid escalation in alcohol consumption, and the high prevalence of alcohol use disorders (AUD) co-occur during the critical period of adolescent brain maturation and refinement. Thus, identifying neurodevelopmental vulnerabilities associated with early and escalating alcohol use, particularly during adolescence, is critical (Casey & Jones, 2010). Adolescence in humans is typically defined as age 9 to 18 years of age (Buchanan, Eccles, & Becker, 1992; Hollenstein & Lougheed, 2013), although more recent neuroimaging studies have begun to redefine the completion of adolescence based on reaching a plateau in structural brain development, which spans the entire second decade and into the early twenties (Bennett & Baird, 2006; Giedd, 2008; Gogtay, et al., 2004), or the period of “emerging adulthood”. The term emerging adulthood has been ascribed to ages 18–25 years (Arnett, 2000), the late adolescent period that overlaps with the final stages of brain maturation (Bennett & Baird, 2006; Gogtay, et al., 2004), in which individuals are functionally more independent than adolescents, but not as independent as individuals aged 25 and older (Arnett, 2001). In order to provide species-specific context for integrating findings across preclinical and clinical studies presented in this review, age spans of adolescence and emerging adulthood as defined in humans roughly correspond in rats to postnatal (P) days 28 to 42 for adolescence (Spear, 2000) and P42 to 55 for late adolescence/“emerging adulthood” (Vetter-O’Hagen & Spear, 2012), in mice from P23 to 35 for pre-adolescence and P36 to 48 for mid adolescence (Adriani, et al., 2004), in cats adolescence is around 11 weeks (Chen, et al., 2007), and in non-human primates adolescence lasts roughly from two to four years (Schwandt, Barr, Suomi, & Higley, 2007).
Adolescent structural and functional maturational brain changes have been well-characterized (Paus, 2005; Spear, 2000), due in part to the increasing availability of non-invasive magnetic resonance (MR) techniques. Consistent with physiological and social maturation from adolescence through emerging adulthood, developmental changes in white matter (WM) and gray matter (GM) tissue volumes have been observed throughout the second decade of life, i.e., adolescence (Giedd, et al., 1999a; Giedd, et al., 1996a; Herting, Maxwell, Irvine, & Nagel, 2012; Nagel, et al., 2006; Pfefferbaum, et al., 1994a; Sowell, Thompson, Tessner, & Toga, 2001; Sowell, Thompson, & Toga, 2004), with WM alterations generally reflecting axon myelination and GM changes reflecting pruning and elimination of synaptic connections (Casey, et al., 2005). Significant longitudinal structural changes, including increased cerebral spinal fluid and GM reductions in frontal and parietal brain regions, have even been detected in as small as a 7-month window during adolescence, suggesting that maturational changes during this age span occur in a very rapid fashion (Sullivan, et al., 2011). Simultaneously, there are dramatic improvements in higher order cognitive abilities (e.g., executive functions) that occur towards the end of the adolescent period and into emerging adulthood (Anderson, 2001; Klenberg, Korkman, & Lahti-Nuuttila, 2001; Rosso, Young, Femia, & Yurgelun-Todd, 2004; Williams, Ponesse, Schachar, Logan, & Tannock, 1999), which are related to a marked, later reorganization and refinement of the frontal lobe and to improved functional WM connectivity within and between brain cortical and subcortical brain regions (Casey, Giedd, & Thomas, 2000; Giedd, et al., 1999b; Pfefferbaum, et al., 1994b; Sowell, Delis, Stiles, & Jernigan, 2001; Sowell, et al., 2004). In addition to these cognitive improvements during adolescence, there is also an increased propensity to seek out novel stimulation and engage in risk-taking behaviors during this time (Arnett, 1992; Trimpop, Kerr, & Kirkcaldy, 1999), likely related to immature cognitive and behavioral response inhibition abilities.
Accordingly, functional magnetic resonance imaging (fMRI) studies document significant increases in the magnitude of frontal lobe activity during the performance of executive function tasks, highlighting the importance of maturation of this region in contributing to age-related improvements in self-regulatory control (Casey, et al., 2000; Dempster, 1992; Durston, et al., 2006; Schweinsburg, et al., 2004; Silveri, Rogowska, McCaffrey, & Yurgelun-Todd, 2011). There has been a paucity of studies utilizing magnetic resonance spectroscopy (MRS) to measure brain chemistry during healthy development (Cohen-Gilbert, Jensen, & Silveri, in press), with only one published study documenting significantly lower GABA levels in the anterior cingulate cortex (ACC) region of the frontal lobe, but not in the parieto-occipital cortex (POC) of adolescents compared to emerging adults (Silveri, et al., 2013) (Figure 2).
Figure 2. Clinical Investigation of In Vivo GABA from Adolescents and Emerging Adults usaig MRS.

Sample raw MRS spectral data were acquired at 4.0 Tesla using MEGAPRESS from the anterior cingulate cortex (ACC, sagittal image insert depicting ACC voxel placement) from an adolescent (orange, left spectrum) and an emerging adult (blue, right spectrum). A smaller ACC GABA peak is evident in the sample adolescent difference-edited MEGAPRESS spectrum compared to the sample from the emerging adult. When the GABA peak was quantified and normalized to creatine for each subject and averaged for each age group, significantly lower ACC brain GABA levels (22.6%, p =.029) were evident in n=30 adolescents (age range: 12–14 yrs, average: 13.6 ± 0.9 (SD) years) compared to n=19 emerging adults (age range: 18–24 yrs, average: 21.6 ± 1.7 (SD) years). See also (Silveri, et al., 2013).
While the frontal cortex undergoes the most substantial structural and functional maturation during adolescence, age-related changes in the medial temporal lobe (specifically, the amygdala and hippocampus) have likewise been reported (Demaster & Ghetti, 2012; Giedd, et al., 1996b; Gogtay, et al., 2006; Sowell & Jernigan, 1998; Suzuki, et al., 2005; Uematsu, et al., 2012). Taken together, adolescent maturational brain changes occurring in important brain regions, the frontal lobe and the hippocampus, continue well into emerging adulthood. Not only are these regions, and their associated functions, the last to reach adult levels (Gogtay, et al., 2004; Gogtay, et al., 2006), these regions are also the first to deteriorate in healthy aging (Raz, et al., 2005; Raz, Rodrigue, & Haacke, 2007), and both are particularly susceptible to the negative consequences of alcohol consumption (Oscar-Berman & Ellis, 1987; Oscar-Berman & Marinkovic, 2007).
On a neural level, refinement of multiple neurotransmitter systems contribute to the overarching structural and functional brain changes observed during adolescence (Spear, 2000), with developmental refinement of the GABA system in particular being linked to changes in cortical-related activity (Cunningham, Bhattacharyya, & Benes, 2008; O’Donnell, 2012; Tseng & O’Donnell, 2007; Vincent, Pabreza, & Benes, 1995). Such GABA changes include enhanced inhibition, which on a behavioral level likely serves to improve self-regulatory control. The following sections describe an abundant preclinical literature across several species (rats, mice, cats, and monkeys) that is continuing to expand regarding the maturation of this main inhibitory neurotransmitter in the mammalian brain (McCormick, 1989). Indeed, developmental maturational changes In GABA occur across many cellular domains, including changes in GABA concentration, synthesis, turnover, and transport, density of GABA receptors, GABA receptor subunit compositions, and GABAergic post synaptic responses (for review, see Kilb, 2012), as described below.
At birth, on a macroscopic level, whole brain GABA concentrations in rodents are approximately 50% of adult levels (Coyle & Enna, 1976). Levels of glutamic acid decarboxylase (GAD), responsible for synthesis of GABA via decarboxylation of glutamate to GABA and carbon dioxide, increase at a slower rate ontogenetically than GABA concentrations (Coyle & Enna, 1976; Johnston & Coyle, 1981). There are two isoforms of GAD, GAD67 mainly catalyzes cytoplasmic GABA pools and GAD65 predominantly catalyzes synaptically directed GABA (Soghomonian & Martin, 1998). In cat visual cortex, levels of GAD67 are detectable at birth and reach adult levels early in life, whereas maturation of GAD65 continues into adolescence (Guo, Kaplan, Cooper, & Mower, 1997). To the extent that GABA plays a number of prominent roles in addition to inhibitory neurotransmission, including metabolism of glucose and fatty acids, GABA concentrations are likely a liberal index of GABA development as compared to enzymatic activity or density of receptors (Coyle & Enna, 1976).
In presynaptic endings, vesicular GABA transporters (VGAT) do not exhibit major developmental changes, whereas plasmalemmal GABA transporters (GAT-1,2,3) located in neurons, glial cells and leptomeningeal cells are typically low at birth in neonatal rat and postmortem human brain, become overexpressed during adolescence and then decline to adult levels (GAT-1) (Xia, Poosch, Whitty, Kapatos, & Bannon, 1993) or are high at birth and up-regulated to reach adult levels during adolescence in rat cerebral cortex (GAT-2,3) (Minelli, Barbaresi, & Conti, 2003). Levels of GAT-1 influence presynaptic uptake and shorten GABAergic postsynaptic potentials, however it is noteworthy that GABA-degrading enzymes, e.g., GABA-transminase (GABA-T), increase to reach adult levels in rats during adolescence (Kristt & Waldman, 1986; Pitts & Quick, 1967), suggesting that adolescent-related modulation of GABA transporters and turnover both contribute to the maturation of GABA neurotransmission.
In the cortex, there are three classes of GABA interneurons, which are categorized based on morphological characteristics, basket (nest, small and large) cells, Chandelier, and Martinotti cells, whereas in the hippocampus, 21 classes of GABA interneurons have been identified (Kilb, 2012). Across these various interneuron cell types, GABAergic interneurons throughout the neocortex undergo massive modifications, in which some interneurons increase linearly to reach adult density after adolescence in monkeys (large basket cells, Erickson & Lewis, 2002), whereas other types of interneurons increase dramatically in density, are transiently overexpressed during adolescence, and are then pruned to adult levels (Chandelier and Martinotti cells) (in monkeys, Anderson, Classey, Conde, Lund, & Lewis, 1995; Cruz, Eggan, & Lewis, 2003; in cats, Wahle, 1993). In a human post mortem study, mRNA expression of seven calcium binding proteins and neuropeptides expressed by GABAergic interneurons were examined in the dorsolateral PFC (DLPFC) of individuals from age 6 weeks to 49 years (Fung, et al., 2010). Expression of parvalbumin (Chandelier and basket cells) and cholecystokinin (large basket cells) demonstrated a delayed increase, with levels peaking during the adolescent years, whereas other calcium binding proteins or neuropeptide expression either peaked earlier in development or decreased slowly over the age span examined (in humans, Fung, et al., 2010; in rats, Kawaguchi & Kondo, 2002). In contrast, GABA interneurons in the hippocampus generally have been found to obtain adult morphological characteristics early in life, prior to adolescence (in mice, Jiang, Oliva, Lam, & Swann, 2001). These developmental changes in GABA interneurons, some of which span, peak or end prior to adolescence, suggest that inhibitory input undergoes considerable refinement prior to adulthood, which can not only influence age-typical improvements in behavior (reduced impulsivity) and cognition (executive functioning, including response inhibition), but also mediate age differences in the responsivity to alcohol effects. However, ontogenetic changes in the GABAergic system that coincide with the adolescent period are considerably more complex than the contributions of GABA interneuron receptor density alone.
Synaptic effects of GABA are mediated through two major receptor subtypes, GABAA and GABAB receptors. GABAA receptors are ionotropic heteropentameric membrane proteins that are assembled from 19 receptor subunits: α1-6, β1-6, γ1-3, δ, ε, θ, π, and ρ1-3 (Olsen & Sieghart, 2008; Sieghart & Sperk, 2002). GABAB receptors are metabotropic heterodimers that consist of two receptor subunits: GABR1 and GABR2. GABAA receptors rise sharply though the early adolescent period in rats (Candy & Martin, 1979), with expression and function of GABAA receptors differing more dramatically during development than GABAB receptors (Henschel, Gipson, & Bordey, 2008). In non-human primates, GABAA receptor density is transiently up-regulated, reaches peak levels during the onset of adolescence and then slowly declines to reach adult levels (Lidow, Goldman-Rakic, & Rakic, 1991). The ontogeny of GABAA binding is more pronounced in humans, because the density of GABAergic receptors increases fivefold around birth and then an additional 100% increase that lasts for the following several weeks (Brooksbank, Atkinson, & Balazs, 1981). Furthermore, data from human studies have shown that while GABAA receptors in subcortical structures (e.g., basal ganglia) reach adult-like levels early during adolescence (~14 years), GABAA receptors reach adult-like levels by 18 years of age in the frontal cortex and by 19.5 years of age in the PFC (Chugani, et al., 2001). While there is evidence for developmental changes in GABAB receptors, rodent studies suggest that GABAB receptors are functional within the first three postnatal weeks, with little evidence for subsequent changes into adulthood (Kilb, 2012). Given that GABAB receptors likely undergo minimal developmental changes during adolescence, this review focuses on the putative role of GABAA receptors in the observed adolescent-specific sensitivity to alcohol. Although it is of note that a growing number of studies provide evidence for a role of the GABAB receptor in regulating alcohol sensitivity via GABAergic synapses (Ariwodola & Weiner, 2004; Wu, Poelchen, & Proctor, 2005), and that treatment with the GABAB receptor agonist baclofen has shown some promise in the treatment of alcoholism in humans (Addolorato, Leggio, Cardone, Ferrulli, & Gasbarrini, 2009; Colombo, et al., 2004).
There are several binding sites on the GABAA receptor, including binding sites for endogenous GABA and exogenous GABA agonists and antagonists, muscimol, gaboxadol, and bicuculline (Henschel, et al., 2008; Sigel & Steinmann, 2012). The GABAA receptor also contains several allosteric binding sites, which provide indirect modulation of the receptor. The following positive allosteric modulators indirectly promote GABA action: benzodiazepines, nonbenzodiazepines (e.g., zolpidem), barbiturates, ethanol, some neuroactive steroids, e.g., allopregnanolone, progesterone derivatives 3α-hydroxy-5α-pregnan-20-one (3α,5α-THP) and 3α,21-dihydroxy-5α-pregnan-20-one (THDOC), and anaesthetics (e.g., propofol) (Sigel & Steinmann, 2012). The following negative allosteric modulators indirectly reduce GABA action: flumazenil, Ro15-4513 and some neurosteroids (e.g., dehydroepiandrosterone (DHEA) and DHEA sulfate, pregnenolone and pregnenolone sulfate). GABA neurotransmission can also be inhibited by the non-competitive antagonist picrotoxin, which binds to or near the pore of the GABA receptor and directly blocks chloride (Cl−) from flowing through the channel, which would subsequently lead to depolarization of the cell membrane (Henschel, et al., 2008; Sigel & Steinmann, 2012).
The binding affinity of GABAA receptor sites, channel conductance and other physiological properties of the GABAA receptor are determined by composition of receptor subunits. While there is a multitude of possible isoforms than can result from the 19 available GABAA receptor subunits, only a restricted number of subunit combinations are observed in brain (Farrant & Kaila, 2007). The minimal requirements to produce a GABA-gated ion channel are inclusion of α and β subunits, and an additional γ or δ subunit. The α1β2γ2 combination is the most abundant, followed by α2β3γ2 and α3β3γ2, and in some GABAA receptors, the γ subunit is replaced by δ, ε, or π, and the β subunit is replaced by θ (Henschel, et al., 2008). GABAA receptors are localized to both postsynaptic and extrasynaptic sites. Postsynaptic receptors mediate phasic inhibition via the release of GABA, an opening of GABAA Cl− channels, and ultimately, hyperpolarization of a depolarized membrane. In contrast, extrasynaptic receptors mediate tonic inhibition via responsiveness to low ambient GABA levels that increase the length of time the GABAA receptor is open (Sigel & Steinmann, 2012). The δ subunit in particular is only found on extrasynaptic GABAA receptors (Farrant & Nusser, 2005; Kilb, 2012). Importantly, it is the composition and arrangement of subunits that determines the functional and pharmacological diversity of GABAA receptors (Minier & Sigel, 2004; Olsen & Sieghart, 2009), including responsiveness to the presence of alcohol (Faingold, N’Gouemo, & Riaz, 1998; Kumar, et al., 2009).
From a developmental perspective, GABAA receptor subunits undergo significant spatial and temporal maturational changes, supporting that expression and function of GABAA receptors during development that differs from adults (Henschel, et al., 2008; Kilb, 2012). The most notable developmental changes in GABAA receptor subunits have been reported for α1, α2, α5, β2, γ2, and δ subunits. In rodent and human studies, expression and regional distribution of the α1 subunit is low at birth but is up-regulated during the early postnatal period, peaking from P21–28 and then declining by 20% to reach adult levels at P60 (Gambarana, Beattie, Rodriguez, & Siegel, 1991; Santerre, Gigante, Landin, & Werner, 2013). In contrast, the α2 subunit is more widely distributed at birth and is either constant or declines until adulthood (Laurie, Wisden, & Seeburg, 1992), although α2 subunits are replaced by α1 subunits (Fritschy, Paysan, Enna, & Mohler, 1994). Levels of α3 are highly expressed in frontal and perirhinal cortex, and are up-regulated during from P10 to P30 (Yu, Wang, Fritschy, Witte, & Redecker, 2006), whereas the α5 subunit demonstrates a pronounced down-regulation from P30-P60 (Laurie, et al., 1992). While β1 has a limited distribution throughout the brain and is constant throughout development, β2 and β3 are more widely distributed and increase predominantly from P14-P21, continuing to increase to reach adult levels by P60 (Gambarana, et al., 1991). In contrast, the γ2 subunit peaks from P14-P21, but then decreases to reach adult levels by P60 (Gambarana, et al., 1991). Lower δ subunit expression in cortex has also recently been reported in adolescents (P28–42) relative to adults (P75) (Santerre, et al., 2013). There are important developmental patterns of subunit expression that are regionally-dependent, such as in the hippocampus, where developmental expression patterns are opposite from those observed in cortex, e.g., in hippocampus α1 subunits are down-regulated, and α2 and α5 subunits are up-regulated (Laurie, et al., 1992; Yu, et al., 2006). In the cerebellum, α1 subunits are also down-regulated, as opposed to the pattern of up-regulation observed in cortex. Similarly, γ2, β2, β3 and δ subunits in the cerebellum are down-regulated throughout development, whereas α6 subunits, which are only expressed in cerebellum, are up-regulated into adulthood (Gutierrez, et al., 1997). Given that the composition and arrangement of GABAA receptor subunits determines responsiveness to drugs, including alcohol, developmental changes in GABAA receptor subunits are a putative mechanism underlying the unique responsiveness of adolescents to alcohol effects relative to adults.
In contrast, the glutamate system, the main excitatory neurotransmitter in brain, and the glutamate type receptor, N-methyl-D-aspartate (NMDA), undergo a somewhat opposite profile of maturational changes during ontogeny. Unlike the slower maturation of GABA receptor density, changes in enzymatic activity, and receptor subunit composition, data from previous rodent studies have shown that there is a transient peak in glutamate activity around 2–3 weeks postnatally (Saransaari & Oja, 1995), and that the density of NMDA receptor sites increases dramatically from postnatal day (P)14 until P21 in rats, and then decreasing to reach adult levels (Pruss, 1993). The transient over-expression of NMDA receptors in a number of brain regions early in ontogeny may be critical for synaptic plasticity, as well as related to an increased susceptibility of young animals to NMDA neurotoxicity (McDonald & Johnston, 1990). A trend for higher in vivo ACC glutamate levels was also observed in human adolescents compared to emerging adults, measured using proton MRS (Silveri, et al., 2013).
Although beyond the scope of this review, neurotransmitter systems other than GABA and glutamate undergo significant developmental changes during adolescence that likely influence responsiveness to alcohol and other drugs of abuse, including but not limited to the dopamine system (Andersen, Thompson, Krenzel, & Teicher, 2002; Brenhouse & Andersen, 2011; Brenhouse, Sonntag, & Andersen, 2008; Chen, et al., 2010). To date, non-invasive MRS methods for detecting in vivo neurotransmitter levels are limited to GABA and glutamate, but as neuroimaging methods continue to evolve, perhaps one-day non-invasive detection of in vivo dopamine and other neurotransmitter levels will be possible. Nonetheless, longitudinal in vivo studies of neurochemical changes in healthy adolescents are needed to characterize the maturational profiles of GABA and glutamate throughout the course of adolescent brain development, as such data are sorely lacking, along with neurotransmitter-related improvements in cognitive functioning and changes in other behavioral features of adolescence. Furthermore, determining developmental profiles of these opposing neural systems would significantly advance our understanding of neurochemical correlates associated with underage alcohol consumption (Silveri, 2012).
2.3. Brain Alterations Associated with Alcohol Use Disorders
Heavy alcohol consumption has been associated with deficits across several domains of cognition in adults (Oscar-Berman, 1990; Oscar-Berman, 2000; Oscar-Berman & Marinkovic, 2003; Parsons & Nixon, 1998), with executive functioning and memory domains most vulnerable to disruptions by alcohol (Oscar-Berman & Ellis, 1987; Oscar-Berman & Marinkovic, 2007). Cognitive impairments have been demonstrated under conditions of acute alcohol challenges, in non-drinkers and drinkers, and in populations of binge drinkers, chronic heavy drinkers, alcohol dependent and recently abstinent alcohol dependent individuals (Fillmore, Marczinski, & Bowman, 2005; Goudriaan, Grekin, & Sher, 2007; Weissenborn & Duka, 2003). MRI and fMRI studies have revealed alterations in brain structure (Harris, et al., 2008a; Pfefferbaum, Sullivan, Mathalon, & Lim, 1997; Sullivan & Pfefferbaum, 2005) and brain activation differences during task performance (Braus, et al., 2001; Grusser, et al., 2004; Heinz, et al., 2007; Marinkovic, et al., 2009; Myrick, et al., 2004; Paulus, et al., 2012; Schuckit, et al., 2012; Tapert, Brown, Baratta, & Brown, 2004a; Tapert, et al., 2001; Trim, et al., 2010), which likely contribute to alcohol-related cognitive deficits. For a comprehensive review of studies investigating adult alcohol dependent populations using a wide variety of non-invasive (MR) and invasive neuroimaging techniques (requiring exposure to ionizing radiation in the form of x-rays or injection with radioactive isotopes, e.g. computed tomography, CT; positron emission tomography, PET; single photon emission tomography, SPECT), see (Buhler & Mann, 2011).
A growing body of literature has extended cognitive and neurobiological alterations observed in adult alcohol-using, abusing and dependent populations to now include alterations in brain structure and function, particularly in frontal networks and hippocampus, in adolescent and young adults with alcohol use disorders (AUDs) (De Bellis, et al., 2000; Tapert, Pulido, Paulus, Schuckit, & Burke, 2004b; Tapert, et al., 2004c) and in adolescent binge drinkers (McQueeny, et al., 2009; Schweinsburg, McQueeny, Nagel, Eyler, & Tapert). In an effort to address whether such alterations observed in adolescent with AUDs are antecedent to or consequences of alcohol use during this period of rapid brain development, studies have been conducted that examine adolescents who have not yet begun to drink or who have minimal alcohol use histories, but who are at risk for developing alcohol abuse problems, based on family history of alcoholism. Results from studies of youth that are family history positive for alcoholism (FH+) demonstrate significant differences in brain structure, function and cognition, as compared to youth with a negative family history (FH−). These findings suggest evidence for a unique genetic contribution that manifests prior to the initiation of alcohol use, and which may serve as a neurobiological vulnerability that confers higher risk for later alcohol abuse or dependence in FH+ youth (Hill, 2004; Hill, et al., 2007; Poon, Ellis, Fitzgerald, & Zucker, 2000; Schweinsburg, et al., 2004; Silveri, Tzilos, Pimentel, & Yurgelun-Todd, 2004; Silveri, Tzilos, & Yurgelun-Todd, 2008; Spadoni, Norman, Schweinsburg, & Tapert, 2008). An alternative, more rigorous approach is to use a prospective study design, in which adolescents who are alcohol naïve are enrolled at baseline and then followed over multiple, subsequent years. A significant proportion of adolescents initiate alcohol consumption at some point during the follow-up period, typically at rates that are consistent with epidemiological use data. In a study by Squeglia and colleagues (Squeglia, et al., 2012), alcohol naïve adolescents at baseline who later transitioned into heavy drinking exhibited reduced fMRI responses (hypoactivation) during performance of a frontally-mediated working memory task at the initial visit, followed by hyperactivation in the same regions at follow-up. These findings suggest that patterns of brain activation observed prior to the onset of drinking may serve as a risk factor for future heavy alcohol use. In a second study, adolescents who had limited substance use at baseline but then transitioned to heavy alcohol use exhibited hypoactivation in frontal, parietal and left cingulate regions during inhibitory processing compared with similarly-aged subjects who remained substance-use free (Norman, et al., 2011). Taken together these valuable longitudinal results suggest that attenuated baseline brain activity, prior to the onset of alcohol use, may predict future substance use problems.
There have also been several proton MRS investigations conducted that have documented alterations in neurochemistry associated with alcohol abuse and alcoholism (Meyerhoff, Durazzo, & Ende, 2013). However, it should be noted that that the available MRS data have been limited to older alcohol dependent or recently abstinent alcoholic males (typically > 40 years of age, except (Silveri, et al., 2014), average age 22 yrs) and only two reports to date have stratified drinking groups to investigate binge alcohol consumption (Meyerhoff, et al., 2004). Further, few studies have included a sufficient number of female participants to permit investigation of sex differences, and no published studies are available examining adolescent alcohol-using or abusing populations using MRS. Moreover, of the existing alcohol adult MRS studies, only six studies to date have examined in vivo brain GABA levels (Abe, et al., 2012; Behar, et al., 1999; Gomez, et al., 2012; Mason, et al., 2006; Mon, et al., 2012; Silveri, et al., 2014), although this is likely to change given increased accessibility to improved GABA detection and quantification methods (Mescher, Merkle, Kirsch, Garwood, & Gruetter, 1998; Stagg, 2013; Stagg, Bachtiar, & Johansen-Berg, 2011b). Recently, the first study employing MRS to measure in vivo GABA levels following an acute alcohol challenge was published (Gomez, et al., 2012). In a cohort of healthy young adult social drinkers, occipital cortex (OCC) GABA levels decreased significantly (13 ± 8%) following intravenous alcohol infusion, which is consistent with alcohol’s ability to potentiate the GABA system. A decrease in in vivo brain GABA levels in human brain following an IV alcohol challenge may seem counterintuitive given limited evidence that in vivo microdialysis studies in animals demonstrate increased GABA levels following an acute alcohol challenge, although findings are mixed. For instance, acute local ethanol increased GABA dialysate in central amygdala, (Roberto, Madamba, Stouffer, Parsons, & Siggins, 2004), acute peripheral alcohol reduced GABA output in the nucleus accumbens (Piepponen, Kiianmaa, & Ahtee, 2002) and ventral pallidum (Kemppainen, Raivio, Nurmi, & Kiianmaa, 2010), or had no effect in the nucleus accumbens (Heidbreder & De Witte, 1993), and acute oral alcohol dose had no effect in the ventral tegmental area – substantia nigra or ventral pallidum (Cowen, Chen, Jarrott, & Lawrence, 1998). Alcohol effects on GABA levels in animal microdialysis studies are likely influenced by regionally specificity of alcohol effects on the GABAergic system, as well as by route of ethanol administration.
In vivo GABA levels measured using MRS reflect total GABA measured from the entire tissue within a voxel of interest, thus small extracellular increases in GABA that may be significant may not be detectable. Alternatively, Gomez and colleagues interpreted the reduction in GABA levels observed in healthy social drinkers during an IV alcohol challenge as reflecting acute reductions in GAD associated with alcohol infusion (Seilicovich, et al., 1985). Indeed, other drug challenges that positively modulate the GABAA receptor are also associated with reductions in brain GABA levels measured using MRS, including reduced OCC GABA levels following an alprazolam challenge (Goddard, et al., 2004) and reduced thalamic GABA levels following a challenge with zolpidem, a nonbenzodiazepine sedative/hypnotic positive modulator of the GABAA receptor (Licata, et al., 2009). Given that the OCC GABA reduction observed in healthy social drinkers was maintained throughout the course of the 70 minute MRS acquisition, while subjects were clamped at a blood alcohol level of .06 mg/dL, it will be of interest in future studies to examine whether GABA changes associated with an alcohol challenge are similar in heavy drinking or dependent populations, are observed after oral exposure, or demonstrate unique time courses of recovery during the descending limb of the blood alcohol curve in dependent versus non-dependent cohorts. The remaining alcohol MRS studies have examined in vivo GABA levels in active heavy drinkers or recently abstinent alcohol dependent patients. Findings are detailed in Table 1 and are described below.
Table 1.
MRS Studies of GABA and Alcohol
| Subjects | Age | 1H MRS | Results | |
|---|---|---|---|---|
|
| ||||
| Silveri, et al., 2014 | 23 BD (11f) | 22 | 4T SVS | ⇓ ACC GABA : ⇓NP |
| 31 LD (13f) | 22 | ACC, POC | ||
|
| ||||
| Abe, et al., 2012 | 40 AD (3f) | 52 | 4T SVS | no CON/AD differences |
| 28 AD + PSA (2f) | 45 | ACC, DLPFC, POC | ⇓ ACC Glu, NAA, Cr, GABA (trend only) | |
| 16 CON (1f) | 49 | |||
|
| ||||
| Gomez, et al., 2012 | 11 LD (8f) acute IV 6% ALC challenge (BrAC 60mg/dl reached in 20min) | 26 | 4T SVS OCC |
⇓ GABA, NAA/NAAG 5 min post IV challenge |
|
| ||||
| Mon, et al., 2012 | 44 AD (5f) [20 baseline, 36 5 weeks ABST] | 54 | 4T SVS ACC, DLPFC, POC |
⇓ ACC Glu, NAA, Cr, no GABA diff |
| 16 CON (2f) | 49 | GABA normalized from 9 ± 4 to 34 ± 7 days of abstinence | ||
|
| ||||
| Mason, et al., 2006 | 16 RDA (0f) | 39 | 2.1T SVS | Glu+Gln s>ns; ⇑NAA : |
| 8 CON (0f) | 39 | OCC | ⇑ Glu+Gln ABST ⇓ GABA ns |
|
|
| ||||
| Behar, et al., 1999 | 5 RDA (?f) | 46 | 2.1T SVS | ⇓ GABA +homocarnosine : |
| 10 CON (?f) | 35 | OCC | ⇓ NP | |
Abbreviations: AD, alcohol dependent; ABST, abstinence; ALC, alcohol; BD, binge drinker; BrAC, breath alcohol concentration; CON, control; IV, intravenous; LD, light drinker; RDA, recently detoxified alcoholic; F, female; PSA, poly-substance abuse; (?) female n unspecified; SVS, single voxel spectroscopy; T, Tesla; ACC, anterior cingulate cortex; DLPFC, dorsolateral prefrontal cortex; OCC, occipital cortex; POC, parieto-occipital cortex; ⇑, significantly higher; ⇓ significantly lower; ns, not significant; : , related to; ns, non-smoker; s, smoker; >, greater than; Cr, creatine; GABA, gamma amino-butyric acid; Gln, glutamine; Glu, glutamate; NAA, N-acetyl-aspartate; NAAG, N-acetyl-aspartate glutamate; NP, neuropsychological performance
The first study, published in 1999 by Behar and colleagues (Behar, et al., 1999), found that OCC GABA + homocarnosine was lower in five recently detoxified alcohol-dependent patients relative to healthy comparison subjects. Seven years later, Mason and colleagues (Mason, et al., 2006) reported that GABA levels in the OCC did not differ significantly between previously alcohol-dependent patients and healthy subjects. In the Mason study, nonsmoking patients had higher GABA levels than smoking patients, however after one month of abstinence, GABA levels decreased significantly in nonsmokers but did not change in smokers. These first two GABA MRS studies document somewhat mixed findings in recently detoxified alcoholics. Although the investigations were limited to the OCC region, smoking status emerged as an important factor to consider when interpreting alcohol effects on brain GABA levels (Cosgrove, et al., 2011). The next two studies, published more recently (2012 and 2013), examined brain GABA levels in regions other than the OCC: ACC, POC and DLPFC in alcohol-dependent individuals during early abstinence. The study by Mon and colleagues (Mon, et al., 2012) found no differences in brain GABA levels at one month of abstinence in alcohol-dependent individuals compared to drug-free controls. In the study by Abe and colleagues (Abe, et al., 2012), alcohol dependent poly-substance users demonstrated significant metabolic abnormalities in the DLPFC and a trend for lower GABA in the ACC, a pattern that was not observed in patients who were only alcohol dependent. While MRS GABA levels acquired in recently abstinent older adults appear largely normal relative to healthy comparison subjects in these two studies, GABA abnormalities in the ACC trended towards significance in concurrent alcohol and illicit drug users. In contrast, GABA MRS data from our laboratory, acquired from a significantly younger cohort of emerging adults (average age 22 years), demonstrate significantly lower GABA levels in the ACC of binge versus light drinkers, while no differences were observed in a comparison region, the POC (Silveri, et al., 2014). Taken together with ontogenetic changes in ACC GABA levels, it is plausible that the dynamically developing adolescent and late adolescent brain may be differentially sensitive to alcohol effects on brain GABA levels, which in turn could serve to modulate age differences in sensitivity to alcohol effects and to the development of alcohol tolerance. It is difficult to discern, however, whether or not low GABA levels observed in the ACC of emerging adult binge drinkers predate the initiation of alcohol consumption. Nevertheless, findings from studies conducted in recently abstinent, adult chronic alcohol abusers offer hope that metabolite abnormalities, GABA in particular, can recover with drinking cessation. Clearly more work is needed to tease apart the role of age of onset of alcohol use and the contribution of psychiatric co-morbidities, as well as to investigate potential relationships between GABA levels and cognitive functioning, all of which could help identify individuals most vulnerable to relapse.
Proton MRS methods for quantifying in vivo GABA metabolite levels using MRS have undergone significant advances over the past decade that have improved detection and quantification of this important neurotransmitter (Hetherington, Newcomer, & Pan, 1998; Jensen, Frederick Bde, & Renshaw, 2005; Keltner, Wald, Frederick, & Renshaw, 1997; Mekle, et al., 2009; Mescher, et al., 1998; Puts & Edden, 2012; Rothman, Petroff, Behar, & Mattson, 1993). GABA levels measured using MRS likely reflect intracellular rather than intrasynaptic levels, since the majority of the GABA pool exists within GABAergic neurons (Petroff, 2002; Stagg, 2013; Stagg, Bachtiar, & Johansen-Berg, 2011b), although determining such specificity is beyond the capabilities of current in vivo MRS (Stagg, 2013). To date, proton MRS remains the only non-invasive method capable of measuring in vivo brain GABA levels in humans, which is particularly advantageous for use in pediatric samples. That is, the non-invasive nature of MRS, compared to exposing the developing brain to radioactive isotopes required for other types of techniques for assessing in vivo GABA function (e.g., PET and SPECT), offers significant promise for future studies to track brain GABA changes associated with alcohol consumption and AUDs during adolescence.
Receptor imaging technologies that include PET and SPECT have been widely employed to probe the role of GABAA receptor function associated with alcohol dependence in adult populations (Cosgrove, et al., 2011). For instance, GABAA receptor density can be inferred in PET using [11C]flumazenil, and more recently [18F]AH114726 (Rodnick, et al., 2013), GABAA benzodiazepine receptor antagonists, and [123l]iomazenil, a GABAA benzodiazepine receptor inverse agonist. Using these radiotracer methods, reduced GABAA receptor densities have been reported in 1-month, 3-month and 7-month abstinence alcohol dependent patients, and from brain regions that include the medial prefrontal cortex and cerebellum (Abi-Dargham, et al., 1998; Lingford-Hughes, et al., 2000; Lingford-Hughes, et al., 1998; Lingford-Hughes, et al., 2005). Conversely, elevated GABAA receptor density has been reported earlier during the course of abstinence, at 1-week, which normalized to control levels by 4-weeks (Staley, et al., 2005). Taken together, these studies suggest that GABAA receptor density undergoes unique temporal changes during the course of recovery from alcohol dependence. Much like MRS, however, these invasive imaging techniques have their limitations, including exposure to radioactive isotopes and an inability to identify changes in GABAA receptor subunit composition or to differentiate between synaptic or extrasynaptic receptor subtypes. Despite advances in imaging technologies, there have yet to be published data from studies of alcohol dependent patients that combine MRS and PET to examine relationships between GABA concentrations and GABA receptor densities, respectively.
Within the framework of characterizing brain reorganization and rapid improvements in cognition during adolescence, this age span is already associated with a reduced capacity to evaluate and appropriately modulate response inhibition and emotional responses (Luna & Sweeney, 2004; Rubia, et al., 2000; Yurgelun-Todd, 2007), related in part to the rapid maturation of frontal and limbic networks, and perhaps related to a developing GABAergic neurochemical system (Silveri, et al., 2013). Accordingly, there are increased challenges within the context of healthy adolescent decision-making that are associated with immature neurobiological machinery, including making the decision to start drinking. That same decision of when to begin drinking not only influences the escalation of alcohol consumption and the likelihood of developing an alcohol abuse disorder later in life (Brown & Tapert, 2004; Chassin, Flora, & King, 2004; Grant & Dawson, 1997; Hill, Shen, Lowers, & Locke, 2000), but also impacts healthy brain development. Due to ethical considerations against administering alcohol to human youth, to better understand the impact of use on multiple domains of development and functioning, animal studies have proven invaluable in identifying the consequences of alcohol use on the brain and on behavior in adolescents versus adults, under controlled laboratory conditions.
3. Preclinical Evidence of Adolescent-Specific Alcohol Sensitivities
Over the past decade, there have been numerous reviews published, based on a multitude of empirical research articles, detailing a unique responsiveness of adolescents to alcohol effects (Chambers, Taylor, & Potenza, 2003; Chin, Van Skike, & Matthews, 2010; Clark, et al., 2008; Doremus-Fitzwater, Varlinskaya, & Spear, 2010; Ehlers & Criado, 2010; Guerri & Pascual, 2010; Masten, Faden, Zucker, & Spear, 2008; Matthews, 2010; Riggs & Greenberg, 2009; Schramm-Sapyta, Walker, Caster, Levin, & Kuhn, 2009; Spear, 2011b; Spear & Varlinskaya, 2005, 2010; Windle, et al., 2008; Witt, 2010). These comprehensive reviews span basic and clinical studies and a variety of topics including neurodevelopment, neurocircuitry of motivation, impulsivity and addiction, age-specific alcohol sensitivities, reinforcing properties of alcohol (e.g., alcohol preference and self-administration), conditioned taste aversion to alcohol, development of tolerance, alcohol-induced withdrawal, and risk characteristics for alcohol initiation, escalation of use, and expression of alcohol abuse disorders. In summary, findings from an extensive collection of empirical preclinical studies illustrate both reduced (hyposensitivity) and enhanced (hypersensitivity) to alcohol effects in adolescents compared to adults.
In general, adolescents are less responsive (hyposensitive) than adults to sedation and motor impairment, and withdrawal-related anxiety, which typically serve to limit intake in humans. In contrast, greater responsiveness (hypersensitivity) to some of the cognitive impairing effects of alcohol has been reported, albeit data are somewhat mixed. While age-related differences in alcohol tolerance likely play a significant role in ontogenetic differences in alcohol sensitivity (Little, Kuhn, Wilson, & Swartzwelder, 1996; Silveri & Spear, 1998, 2001, 2004; Swartzwelder, Richardson, Markwiese-Foerch, Wilson, & Little, 1998), adolescents are more sensitive to alcohol-induced reinforcement (Pautassi, Myers, Spear, Molina, & Spear, 2008), with alcohol exposure during adolescence leading to an increased predisposition to later consume alcohol (Fabio, Nizhnikov, Spear, & Pautassi, 2013). Thus, hypersensitivity to the positive rewarding effects of alcohol could serve to enhance intake. There is also a sizeable literature regarding ontogenetic differences in alcohol-related effects on social behavior, with adolescents demonstrating both greater and reduced sensitivity to unique dimensions of social facilitation and social inhibition, respectively (Varlinskaya & Spear, 2002, 2004b, 2006a, 2006b, 2007; Varlinskaya, Spear, & Spear, 1999, 2001). For the purposes of this review, discussion of adolescent-specific sensitivities to alcohol effects are limited to domains of sedation/ataxia, anxiety, and cognition.
3.1. Adolescent Hyposensitivity to Alcohol Effects
Numerous studies have been published documenting that adolescents are significantly less sensitive than adults to the sedative/hypnotic and motor-impairing effects of alcohol. For instance, longer durations to lose the righting response (LORR) and quicker durations of time to regain the righting response (RORR) have widely been reported in response to acute or repeated alcohol challenges in adolescents relative to adults (Little, et al., 1996; Silveri & Spear, 1998, 1999, 2001, 2002; Swartzwelder, et al., 1998). Chronic intermittent alcohol (CIE) exposure to high doses of alcohol have been shown to prevent the typical age-related increase in the alcohol-induced LORR, due in part to the development of alcohol tolerance (Matthews, Tinsley, Diaz-Granados, Tokunaga, & Silvers, 2008; Silvers, Tokunaga, Mittleman, & Matthews, 2003). With regard to alcohol-induced motor incoordination, several studies demonstrate that adolescents are less impacted than adult counterparts when administered the same dose of alcohol based on body weight (Hollstedt, Olsson, & Rydberg, 1980; Little, et al., 1996; Ramirez & Spear, 2010; White, et al., 2002a; White, et al., 2002b). This age-related insensitivity is underscored by studies demonstrating that significantly higher doses of alcohol are necessary in adolescents to equate alcohol-induced impairments across ages (Broadwater, Varlinskaya, & Spear, 2011; Silveri & Spear, 2001). Indeed, reduced responsiveness to alcohol-induced motor impairment is associated with high alcohol-seeking behavior, at doses that are reflective of those self-administered by adult alcohol-preferring rats (Bell, et al., 2001), and in adolescent alcohol-preferring and high-alcohol-drinking rats (Rodd, et al., 2004).
Anxiety-like behavior that accompanies alcohol withdrawal is typically significant but transient, lasting for longer durations following repeated intermittent bouts of intoxication (Zhang, Morse, Koob, & Schulteis, 2007). To this end, a relative resistance to the anxiogenic behavioral effects of alcohol has also generally been observed following acute alcohol withdrawal in adolescents. While adults demonstrated anxiety-like behavior during acute alcohol withdrawal, measured using the elevated plus maze, similar anxiety profiles were not observed in adolescents, even when alcohol pharmacokinetics were considered (Doremus, Brunell, Varlinskaya, & Spear, 2003; Varlinskaya & Spear, 2004a). Similarly, acute alcohol withdrawal symptoms developed differentially in adolescent versus adult rats, depending on the anxiety symptoms being assessed (Slawecki & Roth, 2004; Slawecki, Roth, & Gilder, 2006). Anxiety-like behavior lasted up to one week following repeated alcohol withdrawals in adolescents, whereas anxiety-like behavior returned to baseline after 24 hours in adults, suggesting that anxiety-like behavior associated with repeated withdrawals is more pronounced in adolescence than in adulthood (Wills, Knapp, Overstreet, & Breese, 2008, 2009). Finally, adolescent rats also demonstrate reduced sensitivity to seizure susceptibility during alcohol withdrawal (Acheson, Richardson, & Swartzwelder, 1999; Chung, Wang, Wehman, & Rhoads, 2008). In contrast, withdrawal severity did not differ between adolescents and adults when a modified Majchrowicz model of alcohol dependence was used to produce similar blood alcohol levels at an expected peak, despite administration of higher alcohol doses to adolescents (3 g/kg per day more) than adults, which was based on lower intoxication scores observed in the younger relative to the older group. Adolescents were qualitatively and quantitatively similar to adults in terms of withdrawal severity after the 4-day binge exposure period, suggesting that alcohol-related withdrawal behavior is largely dependent on blood alcohol levels, regardless of age (Morris, Kelso, Liput, Marshall, & Nixon, 2010). Taken together, adolescents are typically less sensitive to the physiological effects of alcohol, on measures of sedation, ataxia, acute withdrawal-related anxiety and seizure susceptibility, although there are some exceptions. Reduced sensitivity to these alcohol effects can therefore permit adolescents to consume greater amounts of alcohol on a given occasion compared to adults, perhaps given weak physiological feedback regarding intoxication that typically serves to reduce consumption or to terminate a drinking session.
3.2. Adolescent Hypersensitivity to Alcohol Effects
In contrast to the observed physiological hyposensitivities, there is evidence that adolescents are hypersensitive to alcohol’s effects on cognition. In work by Markeweise and colleagues (Markwiese, Acheson, Levin, Wilson, & Swartzwelder, 1998), alcohol exposure was found to impair spatial memory acquisition in adolescent rats, but not in adult rats, and neither group showed impairment on a nonspatial memory task. Follow-up studies using the classic Morris Water Maze task demonstrate that alcohol-impairing effects on cognition are dose-dependent (Acheson, Ross, & Swartzwelder, 2001), and that while alcohol-induced impairments are quickly reversed after treatment ends in adult animals, impairments in spatial memory persist well beyond the termination of repeated alcohol treatment in adolescents (Sircar & Sircar, 2005). Deficits in acquisition of spatial memory emerge in part due to slower learning associated with alcohol exposure in adolescents relative to adults (Novier, Skike, Chin, Diaz-Granados, & Matthews, 2012). In work examining the effects of four weeks of CIE exposure during adolescence in rats, spatial working memory impairments persisted in young adults even after alcohol exposure had ended (Schulteis, Archer, Tapert, & Frank, 2008). Adolescent binge ethanol exposure was also reported to impair reversal learning on the Morris Water Maze and on the Barnes Maze when mice were when mice were assessed in adulthood (Coleman, He, Lee, Styner, & Crews, 2011; Coleman, Liu, Oguz, Styner, & Crews, 2014) ; Crews, Braun, Hoplight, Switzer, & Knapp, 2000). Structural brain enlargements observed in adult orbitofrontal cortex, cerebellum, thalamus, internal capsule and genu of the corpus callosum were associated with adolescent binge ethanol exposure, and were hypothesized to lead to a lack of behavioral flexibility that may contribute to later impairments in spatial memory (Coleman, Liu, Oguz, Styner, & Crews, 2014). Using an appetitive conditioning paradigm, adolescents, but not adults, show impaired discrimination performance if training was followed by alcohol (Land & Spear, 2004a). In contrast, spatial acquisition in adolescents is not impaired on a less stressful sandbox maze relative to adults (Rajendran & Spear, 2004), or when alcohol exposure precedes learning to freeze to a tone (non-hippocampal dependent) or to a context (hippocampal dependent) (Land & Spear, 2004b).
Importantly, alcohol exposure, particularly CIE during adolescence, has been reported to produce cognitive tolerance to subsequent alcohol-induced spatial memory impairments that dissipate by adulthood (Silvers, et al., 2003; Silvers, et al., 2006). However, studies addressing age-differences in metabolic alcohol tolerance highlight that blood alcohol levels alone do not solely underlie alcohol-induced spatial memory impairments (Van Skike, Novier, Diaz-Granados, & Matthews, 2012), as the adolescent rat brain also demonstrates hypersensitivity to alcohol-induced brain damage when binge alcohol exposure occurs during adolescence compared to adulthood (Crews, Braun, Hoplight, Switzer, & Knapp, 2000). Damage to anterior portions of the piriform and perirhinal cortices (important for memory), as well as frontal cortical olfactory regions, measured using an amino cupric silver stain procedure, was observed after four days of binge ethanol treatment in adolescent rats (P35) (Crews, Braun, Hoplight, Switzer, & Knapp, 2000). As discussed in the subsequent section, GABAergic functional changes associated with adolescent alcohol exposure that lead to altered hippocampal physiology likely contribute to the ontogenetic profile of enhanced cognitive impairments associated with alcohol exposure during adolescence (Slawecki, Betancourt, Cole, & Ehlers, 2001; Tokunaga, Silvers, & Matthews, 2006).
3.3. Role of GABA in Adolescent Alcohol Responsiveness
It is plausible that the differential sensitivities to alcohol action in adolescents contribute to an increased tendency for greater alcohol consumption, subsequently leading to neurobiological adaptation to chronic alcohol exposure. Indeed, unique adolescent maturational profiles of central target sites of alcohol action, e.g., GABA and glutamate (Deitrich, Dunwiddie, Harris, & Erwin, 1989; Harris, Trudell, & Mihic, 2008b; Kumar, et al., 2009), may contribute to the neurobiological underpinnings of some of the adolescent-specific sensitivities observed to alcohol effects.
Importantly, GABAA agonists and NMDA antagonists have been shown to potentiate the hypnotic effects of alcohol in adult rats (Beleslin, Djokanovic, Jovanovic Micic, & Samardzic, 1997; Liljequist & Engel, 1982). To the extent that ontogenetic differences in neural mechanisms might underlie the relative resistance of young animals to alcohol, psychopharmacological manipulations targeting GABA and NMDA receptor systems were found to differentially modulate age differences in alcohol responsiveness in adolescent and adult rats (Silveri & Spear, 2002, 2004). As can be seen in Figure 3, the NMDA antagonist, (+)MK-801, administered prior to a sedative dose of alcohol, significantly increased alcohol-induced sedation at both ages, but maintained the age-related increase in sensitivity typically observed. In contrast, adolescents were considerably more sensitive to pretreatment with the GABAA agonist muscimol, which enhanced alcohol sedation to a greater degree in adolescents than in adults. Although modulation of both systems enhanced sedative effects of alcohol across ages, stimulation of the GABAA receptor was a more effective means of enhancing alcohol sedation in adolescents (Figure 3). Given that ontogenetic changes in sensitivity to alcohol sedation are related to a notable age decline in acute tolerance, acute tolerance observed in adolescents was found to be disrupted by (+)MK-801 but not muscimol (Silveri & Spear, 2004). It is plausible that the observed differences are related to differences in the mechanism by which muscimol and +MK-801 target their receptors. Nonetheless, pharmacological attenuation of the expression of acute tolerance was sufficient but not necessary to delay recovery of righting after alcohol. These data suggest that maturation of the GABA and glutamate systems may uniquely contribute to a notably reduced sensitivity to alcohol sedation during ontogeny. Indeed, not only does repeated alcohol exposure lead to the development of alcohol tolerance, both metabolic and neuronal, the latter of which is likely associated with GABAA receptor down regulation, a combination of reduced GABA concentrations associated with ontogenetic state and alcohol-related changes in expression and responsiveness of GABA receptors could have significant ramifications for the observed behavioral sensitivity to alcohol. For a review of GABA-related alcohol tolerance mechanisms associated with acute and chronic exposure, see work by Kumar and colleagues (Kumar, et al., 2009).
Figure 3. Preclinical Investigation of GABA and Glutamate Manipulations on Alcohol-Induced Sedation.
Each bar represents data from n=12 (collapsed over sex) Sprague-Dawley rodents intraperitoneally (ip) injected on postnatal (P) day 26 (adolescents, left set of three bars) or P70 (adults, right set of three bars) with 0.9% saline vehicle (yellow bars), 0.10 mg/kg of the NMDA antagonist (+)-5-methyl-10,11-dihyrdro-5H-dibenzo[a,d]-cyclo-hepten-5,10-imine hydrogen maleate [(+)MK-801] (red bars), or 1.25 mg/kg of the GABA agonist muscimol (green bars). Ten minutes after pretreatment administrations, all animals received an ip injection of 3.5 g/kg of ethanol (17% v/v; diluted with 0.9% saline). Alcohol-induced sedation was measured as the time to RORR (min) as a function of pretreatment [saline, (+)MK-801, or muscimol] and age. Data are expressed as means ± SEM. See also Silveri and Spear, 2002; 2004.
Adolescents were also reported to be less sensitive to a zolpidem, a positive allosteric modulator of the GABAA receptor, compared to older counterparts (Moy, Duncan, Knapp, & Breese, 1998). Measurement of [3H]zolpidem binding in that same study revealed age-related increases in binding, with binding in the cingulate cortex, medial septal nucleus, globus pallidus, inferior colliculus, red nucleus, and cerebellum, increasing up to day P20, stabilizing from P20 through P28, and followed by a subsequent increase into adulthood. Thus, consistent with GABAA and NMDA manipulation studies, adolescent-hyposensitivity to alcohol sedation appears to be coincident with developmental changes in GABAA receptor sites targeted by [3H]zolpidem (Moy, et al., 1998). To the extent that GABA also plays a role in alcohol-induced seizure susceptibility (Olsen & Spigelman, 2012), when pentylenetetrazole was used to induce seizures during alcohol withdrawal, seizures were significantly shorter in adolescents compared to adults (Acheson, et al., 1999). In contrast, modulation of the GABAA receptor with acute alcohol or the neuroactive steroid allopregnanolone did not differentially impair the retrieval of spatial memory on the Morris water maze between adolescents and adults (Chin, et al., 2011).
Recent mounting data from molecular, cellular and behavioral studies implicate perhaps a more notable role of GABAA type receptors in the adaptations to alcohol, including voluntary intake, dependence and withdrawal (Grobin, Matthews, Devaud, & Morrow, 1998). For instance, the GABAA agonist THIP increased intake of and preference for alcohol, while the GABA antagonist picrotoxin decreased intake of and preference for alcohol, over baseline levels in adult rodents (Boyle, Segal, Smith, & Amit, 1993). In adolescents, however, lorazepam, also a GABAA positive modulator, increased 60-minute food intake, whereas the neuroactive steroid DHEA decreased food intake on injection days. In adulthood, lorazepam-treated animals preferred the lowest concentrations of alcohol + saccharin more than saccharin alone, compared with vehicle-treated subjects who showed no preference for any concentration of alcohol + saccharin over saccharin. DHEA-treated animals, however, showed no preference for any of the available solutions. These data demonstrate that GABAA receptor modulation during adolescence can alter intake and preference for ethanol in adulthood, highlighting the importance of drug history as an important variable in the liability for alcohol abuse (Hulin, Amato, & Winsauer, 2012). In contrast, it was reported that exposure to alcohol during adolescence does not necessarily enhance ethanol’s reinforcing properties later in life, findings that could have been related to ethanol initiation/oral self-administration procedures (Tolliver & Samson, 1991). In general, these studies suggest evidence for a diminished potency of alcohol as a positive GABAA modulator, which likely contributes to the reduced sensitivity of adolescents to alcohol effects. Therefore, while alcohol interacts with multiple neurotransmitters systems and receptors, including NMDA, it seems GABA may be one of the most influential in terms of mediating alcohol effects during adolescence.
Although there are mixed results regarding the behavioral effects of alcohol on cognition in adolescent rats, there is a consistent line of evidence documenting enhanced sensitivity to alcohol effects on hippocampal physiology. Hippocampal long-term potentiation (LTP), one of the major cellular mechanisms underlying learning and memory, is attenuated by alcohol exposure. To this end, there are studies documenting hypersensitivity of adolescent hippocampal tissue slices to the inhibitory potency of NMDA receptor-mediated synaptic activity to alcohol (Pyapali, Turner, Wilson, & Swartzwelder, 1999; Swartzwelder, Wilson, & Tayyeb, 1995). Since publication of these seminal ontogenetic alcohol-LTP studies, attention has shifted to examining differential ontogenetic contributions of GABA to alcohol effects on hippocampal physiology. Adolescent hypersensitivity reported for alcohol effects on hippocampal physiology include greater alcohol-induced GABA-mediated inhibitory postsynaptic potentials (IPSPs) in hippocampal CA1 pyramidal cells (Li, Wilson, & Swartzwelder, 2003, 2006) and enhanced extrasynaptic GABAA receptor mediated tonic currents in dentate granule cells (Fleming, Wilson, & Swartzwelder, 2007). Increases in spontaneous firing characteristics and hyperpolarization-activated cation current of hippocampal interneurons were also significantly more pronounced in adolescents than adults, and also persisted in the presence of glutamatergic and GABAergic blockers (Yan, et al., 2009). In contrast, evoked GABAA receptor-mediated inhibitory postsynaptic currents (IPSCs) were less powerfully enhanced by acute alcohol in adolescents than in adults (Li, et al., 2006). Furthermore, when adolescents received CIE, long-lasting changes in the sensitivity of extrasynaptic GABAA receptors to alcohol were observed in dentate gyrus cells in adulthood. Thus, GABAA receptor changes may “lock-in” adolescent hypersensitivity to alcohol in these cells that extend into adulthood (Fleming, Acheson, Moore, Wilson, & Swartzwelder, 2012). Overall, greater alcohol alterations in hippocampal physiology after alcohol exposure is consistent with a heightened sensitivity to alcohol-induced memory impairments in adolescents (Li, Wilson, & Swartzwelder, 2002). Thus, these studies point to the conclusion that developmental differences in the effects of alcohol on interneuron excitation, which likely contributes to age-differences in alcohol effects on cognition, reflect enhanced sensitivity of GABAA receptor-mediated IPSCs in adolescents compared to adults.
While it has been well established that alcohol enhances the function of GABAA receptors (for a comprehensive reviews see (Kumar, et al., 2009; Ueno, et al., 2001), studies are beginning to elucidate the roles of specific receptor subtypes in alcohol-induced behaviors (Boehm, Ponomarev, Blednov, & Harris, 2006; Boehm, et al., 2004; Harris, et al., 2008b; Korpi, et al., 2007; Lobo & Harris, 2008; Wallner, Hanchar, & Olsen, 2006). For instance, evidence from GABA receptor subunit knockout studies demonstrate that some alcohol-related behaviors are dependent on α2-containing GABAA receptors, such as the development of a conditioned taste aversion to alcohol (Blednov, et al., 2013) and alcohol-induced motor stimulant effects (Blednov, et al., 2011), albeit there is some conflicting evidence (Lovinger & Homanics, 2007). It is tempting to categorize subunit involvement by behavioral alterations demonstrated by receptor subunit knock-in and knock-out studies, e.g., altering α1 subunit composition alters sedation, α2/3 alters anxiety and α5 alters temporal and spatial memory, however “behavior is a complex phenomenon, and most probably, there are several types of GABAA receptors involved in even simple behavioral traits” (Sigel & Steinmann, 2012). Nonetheless, enhanced GABAergic function, as it relates to high alcohol preference (Hwang, Kunkler, & Lumeng, 1997), warrants investigation of the potential interplay between ontogenetic changes in GABA subunits and the unique profile of behavioral sensitivity to alcohol effects observed during adolescence.
Finally, aside from alcohol’s ability to act as a GABAA receptor positive modulator, alcohol also increases levels of some neuroactive steroids, including 3α,5α-THP and 3α,5α-THDOC, which have been reported to enhance GABAergic neurotransmission, potentially influencing the behavioral effects of alcohol (Kumar, et al., 2009). Acute alcohol intoxication has also been shown to increase levels of the neuroactive steroid allopregnanolone in the hippocampus of rodents tested during adulthood (P52 and P60), however the alcohol-related increase in allopregnanolone levels was blunted by CIE during adolescence (Silvers, et al., 2006). Likewise, elevated allopregnanolone levels have been reported in human adolescents in association with acute alcohol intoxication (Torres & Ortega, 2003, 2004). Of note, consistent with previous adult GABA MRS data (Epperson, et al., 2006; Epperson, et al., 2002; Epperson, et al., 2005), there is evidence for an influence of menstrual cycle phase, when circulating sex hormones vary, on brain GABA levels measured in both adolescents and emerging adults. Luteal phase (peak estrogen at ovulation followed by declining levels, and high levels of the neuroactive steroid, progesterone) females exhibit lower GABA levels relative to females tested during the follicular phase (low estrogen, low progesterone) (Silveri, et al., 2013). Given the ongoing hormonal changes that accompany brain development and adolescence in general (Blakemore, Burnett, & Dahl, 2010), studies are needed to investigate potential hormone-GABA interactions that could influence alcohol effects during this critical age period. It is plausible that alcohol sensitivity is not only modulated by GABAA receptor activity (Morrow, VanDoren, Penland, & Matthews, 2001), but that adolescent-specific sensitivities associated with an immature GABA system may be driven in part by age differences in levels of circulating neurosteroids, the ability of alcohol to modulate neurosteroid levels, and/or alcohol-neurosteroid interactions at the GABAA receptor.
4. Translating to the Human Condition: In vivo GABA MRS Studies
As discussed in previous sections, in vivo frontal lobe GABA levels, measured using MRS, were lower in healthy adolescents than in healthy emerging adults, yet GABA levels were lower, higher or unaffected in MRS studies of older adults that were recently abstinent from alcohol use. While it would be extremely valuable to conduct an MRS investigation of brain GABA levels in adolescents with alcohol abuse disorders, it is notable that low in vivo GABA levels have been implicated in a number of other psychopathological conditions in adults, including depression, anxiety, epilepsy, cocaine and alcohol dependence (Chang, Cloak, & Ernst, 2003). Moreover, in vivo GABA levels observed in adult human MRS studies significantly predict motor control, anxiety symptoms and cognitive performance, akin to measures that demonstrate differential alcohol effects in ontogenetic animal studies.
Advances in the reliability and versatility of MRS methods available for measuring GABA via MRS provide a valuable step forward (Mullins, et al., 2012), particularly given the high spatial resolution afforded by combining MRI with MRS in order to acquire and quantify GABA levels across multiple brain regions. While it will be important in future studies to examine relationships between brain GABA levels and GABA receptor density in alcohol dependence and following chronic treatments for alcoholism that involve modulation of the GABA system, the following studies offer important insights for understanding the role of in vivo brain concentrations of GABA in various healthy behaviors, as well as in a number of clinical conditions.
4.1. Motor Control and Sleep
It has been reported that alcohol consumption impairs motor ability at blood alcohol levels at 0.06% and higher, including but not limited to alterations in postural control (Modig, Fransson, Magnusson, & Patel, 2012), delayed compensatory muscle responses (Woollacott, 1983), and impaired driving ability (Ferrara, Zancaner, & Giorgetti, 1994; Martin, et al., 2013). A number of MRS studies have been published that confirm an important role of GABA levels in motor control, although none have investigated GABA levels associated with the effects of an acute alcohol challenge on motor control. Nonetheless, developmental increases in GABA from adolescence to emerging adulthood were associated with lower motor impulsivity (Silveri, et al., 2013). In healthy adults, trait urgency in healthy men correlated with DLPFC GABA levels, with higher GABA predicting lower urgency scores (Boy, et al., 2011), and better responsiveness of subconscious motor mechanisms was related to higher GABA levels in the supplementary motor area (Boy, et al., 2010). Higher GABA levels were also associated with slower reaction times (Stagg, Bachtiar, & Johansen-Berg, 2011a) and predicted a better ability to rapidly resolve shifting gaze from one stimulus over another (Sumner, et al., 2010). Finally, higher frontal lobe GABA levels were correlated with lower extroversion personality traits in healthy individuals (Goto, et al., 2010). Across each of these adult studies, higher GABA levels were associated with better motor control. On the contrary, low GABA levels, as is the adolescent phenotype, is associated with worse motor control or with heightened impulsivity. To translate these GABA findings into interpretation for adolescent-specific alcohol sensitivities, adolescents are less sensitive to the motor incoordinating effects of alcohol, presumably due to a diminished potency of alcohol as a positive GABA modulator to produce motor incoordination at this age, unless significantly higher blood alcohol levels are achieved (relative to adults).
In addition to motor-impairing effects, acute alcohol has a soporific effect, which manifests as sedation. While rodent studies assessing alcohol-induced sedation largely reflect loss of a reflexive response rather than sleep per se, disturbed sleep produced by chronic alcohol abuse in humans is predictive of relapse drinking after periods of abstinence (Brower, 2003; Brower, Aldrich, & Hall, 1998; Gann, et al., 2001). Chronic alcohol treatment produces similar sleep-related deficits in rats, including REM suppression, difficulty falling asleep, and difficulty remaining asleep (Mukherjee & Simasko, 2009). Alcohol-preferring rats were also more susceptible to sleep disruptions after binge alcohol exposure, as evidenced by a significant reduction in nonREM sleep and increased wakefulness the day after treatment (Thakkar, Engemann, Sharma, Mohan, & Sahota, 2010). Although no MRS studies to date have investigated sleep disturbances in relation to brain GABA levels in alcohol dependent patients, altered GABA levels have been reported in patients with sleep disorders. Brain GABA levels, acquired from a large brain region including basal ganglia, thalamus, and temporal, parietal, and occipital white matter and cortex, were ~30% lower in patients diagnosed with primary insomnia relative to individuals without sleep complaints (Winkelman, et al., 2008). Furthermore, low GABA levels were negatively correlated with a faster wake after sleep onset on polysomnography measures (Winkelman, et al., 2008). In a subsequent regional investigation, low GABA levels were regionally specific to the ACC and POC, but not the thalamus of patients with primary insomnia (Plante, et al., 2012a). When MRS data were acquired in the evening, however, OCC GABA levels were elevated in primary insomnia patients relative to the comparison group (Morgan, et al., 2012; Plante, Jensen, & Winkelman, 2012b). Although diurnal stability of OCC GABA levels has been documented in healthy adults (Evans, McGonigle, & Edden, 2010), there is evidence of significantly lower frontal lobe GABA levels in an alternate-shift group compared to a day-shift group, highlighting the impact of an irregular work schedule on the maintenance of GABA levels (Kakeda, et al., 2011). Not surprisingly, sleeps aids such as a zolpidem, which modulates the GABAA receptor, decreased GABA levels in the thalamus, but not in the ACC, likely reflecting GABA turnover in healthy adult volunteers (Licata, et al., 2009). Furthermore, administration of propofol, a hypnotic agent used to induce and maintain general anesthesia, significantly increased GABA levels in motor cortex, sensory cortex, hippocampus, thalamus and basal ganglia in healthy young adult volunteers in an unconscious state (Zhang, et al., 2009). Indeed, the unique developmental changes in the sedative effects of alcohol, coupled with changing sleep mechanics during adolescence (Colrain & Baker, 2011) and the impact of alcohol use on sleep quality (Ehlers & Criado, 2010), data from MRS GABA studies of sleep disorders and sedative/hypnotic agents provide interesting potential new directions for adolescent brain research investigating the impact of alcohol use on sleep physiology.
4.2. Anxiety and Seizure Disorders
Anxiety and seizure are associated with alcohol withdrawal, and are present typically in early abstinence. To draw connections to the GABA system, a growing number of MRS investigations have identified that GABA levels are low in patients with anxiety disorders, including individuals with panic disorder, social phobia and obsessive-compulsive disorder (OCD). Patients with panic disorder exhibit lower OCC GABA levels (Goddard, et al., 2001) and an impaired GABA response to an acute benzodiazepine challenge (Goddard, et al., 2004). Low GABA was also reported in the ACC and basal ganglia of medicated subjects with panic disorder (Ham, et al., 2007), although levels were not different in other PFC areas (Hasler, et al., 2009). More recently, evidence of low GABA in the ACC and medial PFC, was reported in a third study of patients with panic disorder. Notably, in that study, GABA differences were more pronounced in individuals with a family history of panic disorder (Long, et al., 2013). Less data are available from patients with other types of anxiety disorders, although thalamic GABA was lower in patients with social anxiety disorder (Pollack, et al., 2008) and medial PFC GABA was lower in patients with OCD (Simpson, et al., 2012). Interestingly, yoga practice, a lifestyle variable, was found to not only enhance GABA levels in healthy adults, but increased GABA was significantly correlated with lower anxiety and depression scores after completing a 12-week yoga regimen (Streeter, et al., 2007; Streeter, et al., 2010). These benefits of yoga are hypothesized to result from increased activity in the parasympathetic nervous system and enhanced GABAergic activity, which could help to ameliorate symptoms across a variety of diseases (Streeter, Gerbarg, Saper, Ciraulo, & Brown, 2012). Finally, ACC GABA levels were positively associated with harm avoidance temperament in healthy subjects, suggesting a potential link between neurochemistry and responsiveness to fear and anxiety (Kim, et al., 2009). Thus, manifestation of abstinence-related anxiety symptoms may be associated with an alcohol-related reduction in GABA levels in adults, whereas in adolescents, alcohol less potently modulates already low GABA levels, thereby minimizing the manifestation of anxiety symptomatology. To the extent that anxiety disorders are often present in adolescents, regardless of alcohol use, future investigations of GABA levels in human adolescents should directly assess relationships with anxiety profiles.
Alcohol abstinence is also associated with increased seizure vulnerability. There is a compelling literature implicating low GABA in epilepsy and documenting successful pharmacological interventions for seizure disorders that are effective via the modulation of GABA (Doelken, et al., 2010; Mueller, et al., 2001; Petroff, Behar, Mattson, & Rothman, 1996a; Petroff, Rothman, Behar, Hyder, & Mattson, 1999; Petroff, Rothman, Behar, & Mattson, 1996b)except (Simister, McLean, Barker, & Duncan, 2003, 2009). Low brain GABA levels were associated with poor seizure control in patients with complex partial seizures (Petroff, et al., 1996b). Furthermore, low GABA levels in the primary somatosensory cortex of patients with chronic epilepsy were significantly correlated with elevated levels of antibodies to GAD, reflecting GABA synthesis (Stagg, et al., 2010). Brain GABA levels, measured using MRS in the OCC, increased significantly following treatment with the antiepileptic drug vigabatrin, which irreversibly inhibits GABA-T. The response of brain GABA levels to vigabatrin not only predicted improved seizure control in epileptic adults (Petroff, et al., 1996a), but also permitted the identification of positive responders (those with reduced seizure activity) to vigabatrin based on lower baseline GABA levels and a more pronounced GABA increase during treatment (Mueller, et al., 2001). MRS measurements of brain GABA have also been used successfully to monitor treatment responses to vigabatrin in children with epilepsy (Novotny, Hyder, Shevell, & Rothman, 1999).
Two studies combining MRS, to measure brain GABA levels, and SPECT or PET, to measure GABAA benzodiazepine receptor binding, have also been published. In adult patients with partial seizures, OCC GABA levels increased three fold after 25–84 days after treatment with vigabatrin, although benzodiazepine binding (measured using [123I]iomazenil] SPECT) did not change (Verhoeff, et al., 1999). Similarly, three days of treatment with vigabatrin increased OCC GABA levels in healthy adult volunteers, however no evidence of GABAA receptor down regulation (measured using 11C-flumazenil (FMZ)-PET) in response to elevated GABA levels was observed (Weber, et al., 1999). Other antiepileptic drugs targeting the GABAergic system used for treating managing seizures have also been shown to elevate GABA levels measured using MRS, in human and rodent tissue. There was a 62% increase in GABA concentration following vigabatrin applied to human neocortical tissue resected during epilepsy surgery, as well as a 13% increase in GABA concentrations following application of gabapentin, an analogue of GABA. Although there was an 82% increase in GABA following vigabatrin applied to healthy tissue from rodents, there were minimal effects of gabapentin on GABA concentrations (non-significant 2% increase in GABA) (Errante, Williamson, Spencer, & Petroff, 2002). Finally, topiramate, a broad spectrum antiepileptic drug that enhances GABAA receptor mediated Cl− flux, among other actions, likewise increased brain GABA in the OCC of epileptic patients, offering partial protection against further seizure activity (Petroff, Hyder, Rothman, & Mattson, 2001).
Given that low GABA levels are thought to convey vulnerability to developing AUDs in adolescents and adults, drugs utilized to increase GABA levels in seizures disorders could be used to test such a hypothesis. To this end, although there have been no published MRS studies examining the effects of GABA promoting agents on GABA levels in alcohol dependent patients, there is growing interest in the use of medications acting on the GABA system for treating alcohol dependence (Caputo & Bernardi, 2010), use of new non-benzodiazepine GABAergic medications for treating alcohol-related withdrawal syndrome (Leggio, Kenna, & Swift, 2008), and in the role of GABAB receptors in addiction in general (Tyacke, Lingford-Hughes, Reed, & Nutt, 2010).
4.3. Cognition
As previously outlined, alcohol exposure (acute and chronic) impairs cognitive functioning across multiple domains (Oscar-Berman, 1990; Oscar-Berman, 2000; Oscar-Berman & Marinkovic, 2003; Parsons & Nixon, 1998). Most of the available studies reporting relationships between MRS-derived measures of GABA and cognitive performance are based on data acquired from healthy populations, although there are a some reports of GABA-cognition correlations in alcohol abusing populations and in patients with schizophrenia. It is unequivocally accepted that cognition abilities rapidly improve during childhood and adolescence, but then slowly deteriorates into senescence. Interestingly, based on the limited data available, in vivo brain GABA levels follow a similar inverted U-shaped curve and cognition, increasing during adolescence (Silveri, et al., 2013) and decreasing into older adulthood (Gao, et al., 2013). Notably, developmental changes in GABA are observed in brain regions (e.g., frontal lobe) that mediate higher order cognitive abilities, such as executive functioning. This domain of cognitive functioning is also particularly vulnerable to chronic alcohol use (Oscar-Berman, 2000; Sullivan & Pfefferbaum, 2005).
The limited number of studies addressing relationships between in vivo brain GABA levels and cognition is related to the recent evolution in the technology required for acquiring GABA metabolite data, i.e., refinement of GABA-optimized MRS sequences, and availability of state-of-the-art head coils and MR scanners. One of the earliest reports documenting a relationship between GABA and cognition found that higher GABA in the primary visual cortex was associated with better performance on a visual discrimination task (Edden, Muthukumaraswamy, Freeman, & Singh, 2009). Healthy subjects who showed greater tactile frequency discrimination also had higher GABA levels in the sensorimotor cortex (Puts, Edden, Evans, McGlone, & McGonigle, 2011). Although cognition was not tested per se, higher ACC resting state GABA concentrations strongly predicted negative fMRI responses while making emotional judgments (Northoff, et al., 2007). Since negative fMRI responses are coupled to decreased neuronal activity, i.e., neuronal inhibition (Pasley, Inglis, & Freeman, 2007; Shmuel, Augath, Oeltermann, & Logothetis, 2006; Shmuel, et al., 2002), these findings provide important evidence linking brain GABA levels with functional brain activation responses during emotional processing. In adolescents, significant correlations were observed between higher ACC GABA levels and better percent accuracy on No-Go inhibition trials (Silveri, et al., 2013). In contrast, ACC GABA did not significantly predict performance on an alternative response inhibition task, the Stroop Color-Naming task (Silveri, et al., 2013). Thus, GABAergic contributions may not necessarily generalize across impulse control indices, particularly when cognitive control tasks require differing sensory or response demands (Stevens, Kiehl, Pearlson, & Calhoun, 2007). Levels of GABA in the POC were not predictive of response inhibition ability, and ACC GABA did not significantly relate to other aspects of cognition, e.g., attention, memory, or general intelligence, suggesting that GABA and cognition relationships have unique regional significance. In contrast, greater decreases in GABA levels in the primary motor cortex following transcranial direct current stimulation (tDCS) were associated with better short-term motor learning, supporting the notion that LTP-dependent learning is dependent on GABA modulation (Stagg, et al., 2011a). In addition, a balance of low GABAergic inhibition and high glutamatergic excitation in the ventromedial PFC of human volunteers predicted better behavioral performance on a reward-guided decision-making task (Jocham, Hunt, Near, & Behrens, 2012).
In a cohort of alcohol-dependent patients who had lower OCC GABA + homocarnosine than healthy comparison subjects, low GABA was significantly associated with worse delayed verbal memory performance (Behar, et al., 1999). GABA levels in ACC, POC or DLPFC, however, did not predict neurocognitive performance in separate cohorts of recently detoxified alcoholic patients (Abe, et al., 2012; Mon, et al., 2012), however, ACC GABA was negatively correlated with auditory verbal learning and POC GABA was positively correlated with visuospatial memory, but only in former alcoholic patients who were also poly-substance users (Abe, et al., 2012). In our study of emerging adults (ages 18–22 years), significantly lower ACC GABA levels predicted worse response inhibition on No Go trials and worse attention/mental flexibility on the Trail Making Test in binge drinkers, a pattern that was not observed for POC GABA levels, or in either region in light drinkers (Silveri, et al., 2014). Finally, in a cohort of patients with schizophrenia, lower GABA in the visual cortex was significantly correlated with worse orientation-specific surround suppression, a behavioral measure of visual inhibition requiring contrast discrimination (Yoon, et al., 2010). On the other hand, unmedicated patients with schizophrenia had elevated medial PFC GABA levels relative to medicated patients and comparison subjects, although there were no significant correlations observed between DLPFC or medial PFC GABA levels and working memory performance (Kegeles, et al., 2012).
Findings from MRS studies examining relationships between GABA levels and cognitive performance are more varied than relationships between GABA, motor measures and anxiety symptoms. In general in healthy individuals, higher GABA levels predict better performance for discrimination or response inhibition tasks, whereas lower GABA levels predict better performance on memory tasks. In patient populations, there are mixed findings of lower GABA and worse memory, inhibition and abstract reasoning, as well as a lack of significant relationships between GABA levels and cognitive performance. Future studies investigating relationships between GABA levels and cognitive impairments associated with active alcohol use are clearly warranted. Furthermore, learning and memory and decision-making are cognitive processes that are crucial elements for the successful navigation through adolescence. Given the vulnerabilities of these processes to alcohol use and the ongoing developmental changes in the GABA system, initiation and escalation of alcohol use during this time could perpetuate an already heightened level of risk-taking, as well as alter the healthy maturation of this very important neural system that could lead to increased risk for developing alcohol abuse disorders.
5. Summary
Over the past two decades, there has been a dramatic increase in knowledge regarding adolescence, through characterization of neurodevelopmental changes captured non-invasively in humans using a variety of MR techniques and from animal studies that have demonstrated adolescence to be a period of unique sensitivity to alcohol effects. Taken together, there are increasing opportunities for translation between findings from animal and human models, in order to gain the most comprehensive understanding of the consequences of alcohol use on the adolescent brain. As discussed in this review, low in vivo brain GABA levels have been reported in healthy adolescents, alcohol abuse disorders, and psychiatric disorders. To this end, in vivo brain GABA studies using MRS offer new possibilities for probing the GABAergic system in both basic and clinical studies, which could provide valuable information that could lead to the development of new interventions for treating alcohol abuse disorders.
Neuroimaging methods capable of capturing dynamic changes in GABAergic function provide an important step forward, particularly in light of GABAA receptor subunit studies that have identified genetic markers associated with increased risk for alcohol dependence. For instance, studies examining variations in single-nucleotide polymorphisms (SNPs) in GABA receptor genes have revealed associations between genetic variations in GABRA1, GABRA2, and GABRG3 and alcohol dependence and behavioral endophenotypes associated with risk for alcoholism, some of which vary by the chromosome investigated. Chromosome 5q GABA receptor gene variations were found to not play a strong role in alcohol dependence (Dick, et al., 2005), but a subsequent study reported that variations in GABRA1 on chromosome 5q were associated with a less severe phenotype of alcohol dependence, history of alcohol-induced blackouts, age of first drunkenness and subjective responsiveness to alcohol (Dick, et al., 2006b). There is ample evidence GABRA2 variations are also associated with alcohol and other substance use disorders (Covault, Gelernter, Hesselbrock, Nellissery, & Kranzler, 2004; Edenberg, et al., 2004; Enoch, 2008; Fehr, et al., 2006; Lappalainen, et al., 2005; Soyka, et al., 2008) except (Matthews, Hoffman, Zezza, Stiffler, & Hill, 2007). More recently, variations within GABRA2 were associated with an attenuated negative response to an oral alcohol challenge, which reflects a reduced sensitivity to the effects of alcohol that is a risk factor vulnerability to developing alcohol use disorders (Uhart, et al., 2013). In a study examining multiple developmental stages, GABRA2 variations were associated with childhood conduct disorder symptoms but not childhood alcohol dependence symptoms. Importantly, evidence for a consistent elevation in risk for alcohol dependence associated with GABRA2 did not manifest until the mid-20s, lasting into adulthood (Dick, et al., 2006a). More recently, adolescent GABRA2 was strongly associated with sensation-seeking behavior, a hallmark of this age period, and the potential pathway by which GABRA2 confers risk for developing alcohol abuse disorders via an early drive to experiment with alcohol (Dick, et al., 2013). GABRG3 variations on chromosome 15q have also been shown to be associated with risk for alcohol dependence (Dick, et al., 2004), as well as variations in genes coding for other neurotransmitters, such as glutamate and serotonin (Edenberg & Foroud, 2006; Hu, et al., 2005; Schumann, et al., 2008). Nonetheless, combination of in vivo GABA MRS studies in combination with genetic assessments of variations in GABA-specific genes will provide a powerful future perspective across multiple levels of analysis.
Finally, pharmacological treatments targeting the GABA system have been efficacious for treating multiple psychiatric conditions (Bhagwagar, et al., 2004; Krystal, et al., 2002; Sanacora, et al., 2003; Sanacora, Mason, Rothman, & Krystal, 2002; Streeter, et al., 2005). Therapies such as cognitive behavioral therapy have a more modest effect on brain GABA levels compared to other pharmacological interventions; however, significant associations between the degree of post-treatment GABA increases and decreases in psychiatric symptoms (Sanacora, et al., 2006) offer promise for similar studies that monitor GABA levels using MRS to identify responsiveness to treatment interventions for alcoholism. Indeed, neurochemical alterations associated with adolescent alcohol abuse and other disorders could inform the development of novel treatment strategies that target GABA, which could enhance functioning and boost mood, and perhaps reduce drinking, via modulation of the GABAergic system. In conclusion, in vivo brain GABA MRS offers a new lens through which the neurophysiology of adolescent-specific sensitivities to alcohol can be viewed, particularly in light of significant developmental change in the GABAergic system.
Acknowledgments
Dr. Jennifer Sneider graciously reviewed multiple drafts of this manuscript, Dr. J. Eric Jensen has been instrumental in the application of GABA MRS to the study of adolescents and emerging adult binge drinkers in the McLean Imaging Center, and Dr. Linda Spear provided invaluable mentoring and scientific inspiration.
This work was supported by K01 AA014651, R01 AA018153 and R21 AA018724 grants (MMS).
Abbreviations
- ACC
anterior cingulate cortex
- ALC
alcohol
- AUD
alcohol use disorder
- CIE
chronic intermittent ethanol
- DLPFC
dorsolateral PFC
- fMRI
functional magnetic resonance imaging
- GABA
gamma amino-butyric acid
- GAD
glutamic acid decarboxylase
- GM
gray matter
- IPSPs
inhibitory postsynaptic potentials
- LORR
lose the righting response
- LTP
long-term potentiation
- MR
magnetic resonance
- MRI
magnetic resonance imaging
- MRS
magnetic resonance spectroscopy
- MTL
medial temporal lobe
- NMDA
N-methyl-D-aspartate
- OCC
occipital cortex
- P
postnatal day
- PET
positron emission tomography
- PFC
prefrontal cortex
- POC
parieto-occipital cortex
- PTZ
pentylenetetrazole
- RORR
righting response
- SPECT
single photon emission computed tomography
- tDCS
transcranial direct current stimulation
- WM
white matter
Footnotes
Conflict of Interest
The author declares that there are no conflicts of interest.
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- Abe C, Mon A, Durazzo TC, Pennington DL, Schmidt TP, Meyerhoff DJ. Polysubstance and alcohol dependence: Unique abnormalities of magnetic resonance-derived brain metabolite levels. Drug Alcohol Depend. 2012 doi: 10.1016/j.drugalcdep.2012.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abi-Dargham A, Krystal JH, Anjilvel S, Scanley BE, Zoghbi S, Baldwin RM, Rajeevan N, Ellis S, Petrakis IL, Seibyl JP, Charney DS, Laruelle M, Innis RB. Alterations of benzodiazepine receptors in type II alcoholic subjects measured with SPECT and [123I]iomazenil. Am J Psychiatry. 1998;155:1550–1555. doi: 10.1176/ajp.155.11.1550. [DOI] [PubMed] [Google Scholar]
- Acheson SK, Richardson R, Swartzwelder HS. Developmental changes in seizure susceptibility during ethanol withdrawal. Alcohol. 1999;18:23–26. doi: 10.1016/s0741-8329(98)00063-9. [DOI] [PubMed] [Google Scholar]
- Acheson SK, Ross EL, Swartzwelder HS. Age-independent and dose-response effects of ethanol on spatial memory in rats. Alcohol. 2001;23:167–175. doi: 10.1016/s0741-8329(01)00127-6. [DOI] [PubMed] [Google Scholar]
- Addolorato G, Leggio L, Cardone S, Ferrulli A, Gasbarrini G. Role of the GABA(B) receptor system in alcoholism and stress: focus on clinical studies and treatment perspectives. Alcohol. 2009;43:559–563. doi: 10.1016/j.alcohol.2009.09.031. [DOI] [PubMed] [Google Scholar]
- Adriani W, Granstrem O, Macri S, Izykenova G, Dambinova S, Laviola G. Behavioral and neurochemical vulnerability during adolescence in mice: studies with nicotine. Neuropsychopharmacology. 2004;29:869–878. doi: 10.1038/sj.npp.1300366. [DOI] [PubMed] [Google Scholar]
- Andersen SL, Thompson AP, Krenzel E, Teicher MH. Pubertal changes in gonadal hormones do not underlie adolescent dopamine receptor overproduction. Psychoneuroendocrinology. 2002;27:683–691. doi: 10.1016/s0306-4530(01)00069-5. [DOI] [PubMed] [Google Scholar]
- Anderson SA, Classey JD, Conde F, Lund JS, Lewis DA. Synchronous development of pyramidal neuron dendritic spines and parvalbumin-immunoreactive chandelier neuron axon terminals in layer III of monkey prefrontal cortex. Neuroscience. 1995;67:7–22. doi: 10.1016/0306-4522(95)00051-j. [DOI] [PubMed] [Google Scholar]
- Anderson V. Assessing executive functions in children: biological, psychological, and developmental considerationst. Pediatr Rehabil. 2001;4:119–136. doi: 10.1080/13638490110091347. [DOI] [PubMed] [Google Scholar]
- Ariwodola OJ, Weiner JL. Ethanol potentiation of GABAergic synaptic transmission may be self-limiting: role of presynaptic GABA(B) receptors. J Neurosci. 2004;24:10679–10686. doi: 10.1523/JNEUROSCI.1768-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arnett JJ. Reckless Behavior In Adolescence - A Developmental Perspective. Dev Rev. 1992;12:339–373. [Google Scholar]
- Arnett JJ. Emerging adulthood. A theory of development from the late teens through the twenties. Am Psychol. 2000;55:469–480. [PubMed] [Google Scholar]
- Arnett JJ. Conceptions of the transition to adulthood: Perspectives from adolescence through midlife. J Adult Dev. 2001;8:133–143. [Google Scholar]
- Bates ME, Labouvie EW. Adolescent risk factors and the prediction of persistent alcohol and drug use into adulthood. Alcohol Clin Exp Res. 1997;21:944–950. [PubMed] [Google Scholar]
- Bava S, Tapert SF. Adolescent brain development and the risk for alcohol and other drug problems. Neuropsychol Rev. 2010;20:398–413. doi: 10.1007/s11065-010-9146-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Behar KL, Rothman DL, Petersen KF, Hooten M, Delaney R, Petroff OA, Shulman GI, Navarro V, Petrakis IL, Charney DS, Krystal JH. Preliminary evidence of low cortical GABA levels in localized 1H-MR spectra of alcohol-dependent and hepatic encephalopathy patients. Am J Psychiatry. 1999;156:952–954. doi: 10.1176/ajp.156.6.952. [DOI] [PubMed] [Google Scholar]
- Beleslin DB, Djokanovic N, Jovanovic Micic D, Samardzic R. Opposite effects of GABAA and NMDA receptor antagonists on ethanol-induced behavioral sleep in rats. Alcohol. 1997;14:167–173. doi: 10.1016/s0741-8329(96)00140-1. [DOI] [PubMed] [Google Scholar]
- Bell RL, Stewart RB, Woods JE, 2nd, Lumeng L, Li TK, Murphy JM, McBride WJ. Responsivity and development of tolerance to the motor impairing effects of moderate doses of ethanol in alcohol-preferring (P) and -nonpreferring (NP) rat lines. Alcohol Clin Exp Res. 2001;25:644–650. [PubMed] [Google Scholar]
- Bennett CM, Baird AA. Anatomical changes in the emerging adult brain: a voxel-based morphometry study. Hum Brain Mapp. 2006;27:766–777. doi: 10.1002/hbm.20218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bhagwagar Z, Wylezinska M, Jezzard P, Evans J, Boorman E, PMM, PJC Low GABA concentrations in occipital cortex and anterior cingulate cortex in medication-free, recovered depressed patients. Int J Neuropsychopharmacol. 2008;11:255–260. doi: 10.1017/S1461145707007924. [DOI] [PubMed] [Google Scholar]
- Bhagwagar Z, Wylezinska M, Taylor M, Jezzard P, Matthews PM, Cowen PJ. Increased brain GABA concentrations following acute administration of a selective serotonin reuptake inhibitor. Am J Psychiatry. 2004;161:368–370. doi: 10.1176/appi.ajp.161.2.368. [DOI] [PubMed] [Google Scholar]
- Blakemore SJ. Imaging brain development: the adolescent brain. Neuroimage. 2012;61:397–406. doi: 10.1016/j.neuroimage.2011.11.080. [DOI] [PubMed] [Google Scholar]
- Blakemore SJ, Burnett S, Dahl RE. The role of puberty in the developing adolescent brain. Hum Brain Mapp. 2010;31:926–933. doi: 10.1002/hbm.21052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blednov YA, Benavidez JM, Black M, Chandra D, Homanics GE, Rudolph U, Harris RA. Linking GABA(A) receptor subunits to alcohol-induced conditioned taste aversion and recovery from acute alcohol intoxication. Neuropharmacology. 2013;67:46–56. doi: 10.1016/j.neuropharm.2012.10.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blednov YA, Borghese CM, McCracken ML, Benavidez JM, Geil CR, Osterndorff-Kahanek E, Werner DF, Iyer S, Swihart A, Harrison NL, Homanics GE, Harris RA. Loss of ethanol conditioned taste aversion and motor stimulation in knockin mice with ethanol-insensitive alpha2-containing GABA(A) receptors. J Pharmacol Exp Ther. 2011;336:145–154. doi: 10.1124/jpet.110.171645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boehm SL, 2nd, Ponomarev I, Blednov YA, Harris RA. From gene to behavior and back again: new perspectives on GABAA receptor subunit selectivity of alcohol actions. Adv Pharmacol. 2006;54:171–203. doi: 10.1016/s1054-3589(06)54008-6. [DOI] [PubMed] [Google Scholar]
- Boehm SL, 2nd, Ponomarev I, Jennings AW, Whiting PJ, Rosahl TW, Garrett EM, Blednov YA, Harris RA. gamma-Aminobutyric acid A receptor subunit mutant mice: new perspectives on alcohol actions. Biochem Pharmacol. 2004;68:1581–1602. doi: 10.1016/j.bcp.2004.07.023. [DOI] [PubMed] [Google Scholar]
- Boy F, Evans CJ, Edden RA, Lawrence AD, Singh KD, Husain M, Sumner P. Dorsolateral prefrontal gamma-aminobutyric acid in men predicts individual differences in rash impulsivity. Biol Psychiatry. 2011;70:866–872. doi: 10.1016/j.biopsych.2011.05.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boy F, Evans CJ, Edden RA, Singh KD, Husain M, Sumner P. Individual differences in subconscious motor control predicted by GABA concentration in SMA. Curr Biol. 2010;20:1779–1785. doi: 10.1016/j.cub.2010.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boyle AE, Segal R, Smith BR, Amit Z. Bidirectional effects of GABAergic agonists and antagonists on maintenance of voluntary ethanol intake in rats. Pharmacol Biochem Behav. 1993;46:179–182. doi: 10.1016/0091-3057(93)90338-t. [DOI] [PubMed] [Google Scholar]
- Braus DF, Wrase J, Grusser S, Hermann D, Ruf M, Flor H, Mann K, Heinz A. Alcohol-associated stimuli activate the ventral striatum in abstinent alcoholics. J Neural Transm. 2001;108:887–894. doi: 10.1007/s007020170038. [DOI] [PubMed] [Google Scholar]
- Brenhouse HC, Andersen SL. Developmental trajectories during adolescence in males and females: a cross-species understanding of underlying brain changes. Neurosci Biobehav Rev. 2011;35:1687–1703. doi: 10.1016/j.neubiorev.2011.04.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brenhouse HC, Sonntag KC, Andersen SL. Transient D1 dopamine receptor expression on prefrontal cortex projection neurons: relationship to enhanced motivational salience of drug cues in adolescence. J Neurosci. 2008;28:2375–2382. doi: 10.1523/JNEUROSCI.5064-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Broadwater M, Varlinskaya EI, Spear LP. Different chronic ethanol exposure regimens in adolescent and adult male rats: effects on tolerance to ethanol-induced motor impairment. Behav Brain Res. 2011;225:358–362. doi: 10.1016/j.bbr.2011.07.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brooksbank BW, Atkinson DJ, Balazs R. Biochemical development of the human brain. II. some parameters of the GABA-ergic system. Dev Neurosci. 1981;4:188–200. doi: 10.1159/000112756. [DOI] [PubMed] [Google Scholar]
- Brower KJ. Insomnia, alcoholism and relapse. Sleep Med Rev. 2003;7:523–539. doi: 10.1016/s1087-0792(03)90005-0. [DOI] [PubMed] [Google Scholar]
- Brower KJ, Aldrich MS, Hall JM. Polysomnographic and subjective sleep predictors of alcoholic relapse. Alcohol Clin Exp Res. 1998;22:1864–1871. [PubMed] [Google Scholar]
- Brown SA, Tapert SF. Adolescence and the trajectory of alcohol use: basic to clinical studies. Ann N Y Acad Sci. 2004;1021:234–244. doi: 10.1196/annals.1308.028. [DOI] [PubMed] [Google Scholar]
- Buchanan CM, Eccles JS, Becker JB. Are adolescents the victims of raging hormones: evidence for activational effects of hormones on moods and behavior at adolescence. Psychol Bull. 1992;111:62–107. doi: 10.1037/0033-2909.111.1.62. [DOI] [PubMed] [Google Scholar]
- Buhler M, Mann K. Alcohol and the human brain: a systematic review of different neuroimaging methods. Alcohol Clin Exp Res. 2011;35:1771–1793. doi: 10.1111/j.1530-0277.2011.01540.x. [DOI] [PubMed] [Google Scholar]
- Candy JM, Martin IL. The postnatal development of the benzodiazepine receptor in the cerebral cortex and cerebellum of rat. J Neurochem. 1979;32:655–658. doi: 10.1111/j.1471-4159.1979.tb00402.x. [DOI] [PubMed] [Google Scholar]
- Caputo F, Bernardi M. Medications acting on the GABA system in the treatment of alcoholic patients. Curr Pharm Des. 2010;16:2118–2125. doi: 10.2174/138161210791516468. [DOI] [PubMed] [Google Scholar]
- Casey BJ, Galvan A, Hare TA. Changes in cerebral functional organization during cognitive development. Curr Opin Neurobiol. 2005;15:239–244. doi: 10.1016/j.conb.2005.03.012. [DOI] [PubMed] [Google Scholar]
- Casey BJ, Giedd JN, Thomas KM. Structural and functional brain development and its relation to cognitive development. Biol Psychol. 2000;54:241–257. doi: 10.1016/s0301-0511(00)00058-2. [DOI] [PubMed] [Google Scholar]
- Casey BJ, Jones RM. Neurobiology of the adolescent brain and behavior: implications for substance use disorders. J Am Acad Child Adolesc Psychiatry. 2010;49:1189–1201. doi: 10.1016/j.jaac.2010.08.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chambers RA, Taylor JR, Potenza MN. Developmental neurocircuitry of motivation in adolescence: a critical period of addiction vulnerability. Am J Psychiatry. 2003;160:1041–1052. doi: 10.1176/appi.ajp.160.6.1041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang L, Cloak CC, Ernst T. Magnetic resonance spectroscopy studies of GABA in neuropsychiatric disorders. J Clin Psychiatry. 2003;64(Suppl 3):7–14. [PubMed] [Google Scholar]
- Chassin L, Flora DB, King KM. Trajectories of alcohol and drug use and dependence from adolescence to adulthood: The effects of familial alcoholism and personality. J Abnorm Psychol. 2004;113:483–498. doi: 10.1037/0021-843X.113.4.483. [DOI] [PubMed] [Google Scholar]
- Chen X, Wilson RM, Kubo H, Berretta RM, Harris DM, Zhang X, Jaleel N, MacDonnell SM, Bearzi C, Tillmanns J, Trofimova I, Hosoda T, Mosna F, Cribbs L, Leri A, Kajstura J, Anversa P, Houser SR. Adolescent feline heart contains a population of small, proliferative ventricular myocytes with immature physiological properties. Circ Res. 2007;100:536–544. doi: 10.1161/01.RES.0000259560.39234.99. [DOI] [PubMed] [Google Scholar]
- Chen YI, Choi JK, Xu H, Ren J, Andersen SL, Jenkins BG. Pharmacologic neuroimaging of the ontogeny of dopamine receptor function. Dev Neurosci. 2010;32:125–138. doi: 10.1159/000286215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chin VS, Van Skike CE, Berry RB, Kirk RE, Diaz-Granados J, Matthews DB. Effect of acute ethanol and acute allopregnanolone on spatial memory in adolescent and adult rats. Alcohol. 2011;45:473–483. doi: 10.1016/j.alcohol.2011.03.001. [DOI] [PubMed] [Google Scholar]
- Chin VS, Van Skike CE, Matthews DB. Effects of ethanol on hippocampal function during adolescence: a look at the past and thoughts on the future. Alcohol. 2010;44:3–14. doi: 10.1016/j.alcohol.2009.10.015. [DOI] [PubMed] [Google Scholar]
- Chugani DC, Muzik O, Juhasz C, Janisse JJ, Ager J, Chugani HT. Postnatal maturation of human GABAA receptors measured with positron emission tomography. Ann Neurol. 2001;49:618–626. [PubMed] [Google Scholar]
- Chung CS, Wang J, Wehman M, Rhoads DE. Severity of alcohol withdrawal symptoms depends on developmental stage of Long-Evans rats. Pharmacol Biochem Behav. 2008;89:137–144. doi: 10.1016/j.pbb.2007.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clark DB, Thatcher DL, Tapert SF. Alcohol, psychological dysregulation, and adolescent brain development. Alcohol Clin Exp Res. 2008;32:375–385. doi: 10.1111/j.1530-0277.2007.00601.x. [DOI] [PubMed] [Google Scholar]
- Cohen-Gilbert JE, Jensen JE, Silveri MM. Contributions of magnetic resonance spectroscopy to understanding development: Potential applications in the study of adolescent alcohol use and abuse. Dev Psychopathol. doi: 10.1017/S0954579414000030. (in press) [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coleman LG, Jr, He J, Lee J, Styner M, Crews FT. Adolescent binge drinking alters adult brain neurotransmitter gene expression, behavior, brain regional volumes, and neurochemistry in mice. Alcohol Clin Exp Res. 2011;35:671–688. doi: 10.1111/j.1530-0277.2010.01385.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coleman LG, Jr, Liu W, Oguz I, Styner M, Crews FT. Adolescent binge ethanol treatment alters adult brain regional volumes, cortical extracellular matrix protein and behavioral flexibility. Pharmacol Biochem Behav. 2014;116:142–151. doi: 10.1016/j.pbb.2013.11.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Colombo G, Addolorato G, Agabio R, Carai MA, Pibiri F, Serra S, Vacca G, Gessa GL. Role of GABA(B) receptor in alcohol dependence: reducing effect of baclofen on alcohol intake and alcohol motivational properties in rats and amelioration of alcohol withdrawal syndrome and alcohol craving in human alcoholics. Neurotox Res. 2004;6:403–414. doi: 10.1007/BF03033315. [DOI] [PubMed] [Google Scholar]
- Colrain IM, Baker FC. Changes in sleep as a function of adolescent development. Neuropsychol Rev. 2011;21:5–21. doi: 10.1007/s11065-010-9155-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cosgrove KP, Esterlis I, Mason GF, Bois F, O’Malley SS, Krystal JH. Neuroimaging insights into the role of cortical GABA systems and the influence of nicotine on the recovery from alcohol dependence. Neuropharmacology. 2011;60:1318–1325. doi: 10.1016/j.neuropharm.2011.01.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Covault J, Gelernter J, Hesselbrock V, Nellissery M, Kranzler HR. Allelic and haplotypic association of GABRA2 with alcohol dependence. Am J Med Genet B Neuropsychiatr Genet. 2004;129B:104–109. doi: 10.1002/ajmg.b.30091. [DOI] [PubMed] [Google Scholar]
- Cowen M, Chen F, Jarrott B, Lawrence AJ. Effects of acute ethanol on GABA release and GABA(A) receptor density in the rat mesolimbic system. Pharmacol Biochem Behav. 1998;59:51–57. doi: 10.1016/s0091-3057(97)00390-0. [DOI] [PubMed] [Google Scholar]
- Coyle JT, Enna SJ. Neurochemical aspects of the ontogenesis of GABAnergic neurons in the rat brain. Brain Res. 1976;111:119–133. doi: 10.1016/0006-8993(76)91053-2. [DOI] [PubMed] [Google Scholar]
- Crews FT, Braun CJ, Hoplight B, Switzer RC, 3rd, Knapp DJ. Binge ethanol consumption causes differential brain damage in young adolescent rats compared with adult rats. Alcohol Clin Exp Res. 2000;24:1712–1723. [PubMed] [Google Scholar]
- Cruz DA, Eggan SM, Lewis DA. Postnatal development of pre- and postsynaptic GABA markers at chandelier cell connections with pyramidal neurons in monkey prefrontal cortex. J Comp Neurol. 2003;465:385–400. doi: 10.1002/cne.10833. [DOI] [PubMed] [Google Scholar]
- Cunningham MG, Bhattacharyya S, Benes FM. Increasing Interaction of amygdalar afferents with GABAergic interneurons between birth and adulthood. Cereb Cortex. 2008;18:1529–1535. doi: 10.1093/cercor/bhm183. [DOI] [PubMed] [Google Scholar]
- De Bellis MD, Clark DB, Beers SR, Soloff PH, Boring AM, Hall J, Kersh A, Keshavan MS. Hippocampal volume in adolescent-onset alcohol use disorders. Am J Psychiatry. 2000;157:737–744. doi: 10.1176/appi.ajp.157.5.737. [DOI] [PubMed] [Google Scholar]
- Deitrich RA, Dunwiddie TV, Harris RA, Erwin VG. Mechanism of action of ethanol: initial central nervous system actions. Pharmacol Rev. 1989;41:489–537. [PubMed] [Google Scholar]
- Demaster DM, Ghetti S. Developmental differences in hippocampal and cortical contributions to episodic retrieval. Cortex. 2012 doi: 10.1016/j.cortex.2012.08.004. [DOI] [PubMed] [Google Scholar]
- Dempster FN. The rise and fall of the inhibitory mechanism - Toward a unified theory of cognitive-development and aging. Dev Rev. 1992;12:45–75. [Google Scholar]
- Dick DM, Aliev F, Latendresse S, Porjesz B, Schuckit M, Rangaswamy M, Hesselbrock V, Edenberg H, Nurnberger J, Agrawal A, Bierut L, Wang J, Bucholz K, Kuperman S, Kramer J. How Phenotype and Developmental Stage Affect the Genes We Find: GABRA2 and Impulsivity. Twin Res Hum Genet. 2013:1–9. doi: 10.1017/thg.2013.20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dick DM, Bierut L, Hinrichs A, Fox L, Bucholz KK, Kramer J, Kuperman S, Hesselbrock V, Schuckit M, Almasy L, Tischfield J, Porjesz B, Begleiter H, Nurnberger J, Jr, Xuei X, Edenberg HJ, Foroud T. The role of GABRA2 in risk for conduct disorder and alcohol and drug dependence across developmental stages. Behav Genet. 2006a;36:577–590. doi: 10.1007/s10519-005-9041-8. [DOI] [PubMed] [Google Scholar]
- Dick DM, Edenberg HJ, Xuei X, Goate A, Hesselbrock V, Schuckit M, Crowe R, Foroud T. No association of the GABAA receptor genes on chromosome 5 with alcoholism in the collaborative study on the genetics of alcoholism sample. Am J Med Genet B Neuropsychiatr Genet. 2005;132B:24–28. doi: 10.1002/ajmg.b.30058. [DOI] [PubMed] [Google Scholar]
- Dick DM, Edenberg HJ, Xuei X, Goate A, Kuperman S, Schuckit M, Crowe R, Smith TL, Porjesz B, Begleiter H, Foroud T. Association of GABRG3 with alcohol dependence. Alcohol Clin Exp Res. 2004;28:4–9. doi: 10.1097/01.ALC.0000108645.54345.98. [DOI] [PubMed] [Google Scholar]
- Dick DM, Plunkett J, Wetherill LF, Xuei X, Goate A, Hesselbrock V, Schuckit M, Crowe R, Edenberg HJ, Foroud T. Association between GABRA1 and drinking behaviors in the collaborative study on the genetics of alcoholism sample. Alcohol Clin Exp Res. 2006b;30:1101–1110. doi: 10.1111/j.1530-0277.2006.00136.x. [DOI] [PubMed] [Google Scholar]
- Doelken MT, Hammen T, Bogner W, Mennecke A, Stadlbauer A, Boettcher U, Doerfler A, Stefan H. Alterations of intracerebral gamma-aminobutyric acid (GABA) levels by titration with levetiracetam in patients with focal epilepsies. Epilepsia. 2010;51:1477–1482. doi: 10.1111/j.1528-1167.2010.02544.x. [DOI] [PubMed] [Google Scholar]
- Doremus TL, Brunell SC, Varlinskaya EI, Spear LP. Anxiogenic effects during withdrawal from acute ethanol in adolescent and adult rats. Pharmacol Biochem Behav. 2003;75:411–418. doi: 10.1016/s0091-3057(03)00134-5. [DOI] [PubMed] [Google Scholar]
- Doremus-Fitzwater TL, Varlinskaya EI, Spear LP. Motivational systems in adolescence: possible implications for age differences in substance abuse and other risk-taking behaviors. Brain Cogn. 2010;72:114–123. doi: 10.1016/j.bandc.2009.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Durston S, Davidson MC, Tottenham N, Galvan A, Spicer J, Fossella JA, Casey BJ. A shift from diffuse to focal cortical activity with development. Developmental Science. 2006;9:1–20. doi: 10.1111/j.1467-7687.2005.00454.x. [DOI] [PubMed] [Google Scholar]
- Edden RA, Muthukumaraswamy SD, Freeman TC, Singh KD. Orientation discrimination performance is predicted by GABA concentration and gamma oscillation frequency in human primary visual cortex. J Neurosci. 2009;29:15721–15726. doi: 10.1523/JNEUROSCI.4426-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Edenberg HJ, Dick DM, Xuei X, Tian H, Almasy L, Bauer LO, Crowe RR, Goate A, Hesselbrock V, Jones K, Kwon J, Li TK, Nurnberger JI, Jr, O’Connor SJ, Reich T, Rice J, Schuckit MA, Porjesz B, Foroud T, Begleiter H. Variations in GABRA2, encoding the alpha 2 subunit of the GABA(A) receptor, are associated with alcohol dependence and with brain oscillations. Am J Hum Genet. 2004;74:705–714. doi: 10.1086/383283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Edenberg HJ, Foroud T. The genetics of alcoholism: identifying specific genes through family studies. Addict Biol. 2006;11:386–396. doi: 10.1111/j.1369-1600.2006.00035.x. [DOI] [PubMed] [Google Scholar]
- Ehlers CL, Criado JR. Adolescent ethanol exposure: does it produce long-lasting electrophysiological effects? Alcohol. 2010;44:27–37. doi: 10.1016/j.alcohol.2009.09.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Enoch MA. The role of GABA(A) receptors in the development of alcoholism. Pharmacol Biochem Behav. 2008;90:95–104. doi: 10.1016/j.pbb.2008.03.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Epperson CN, Gueorguieva R, Czarkowski KA, Stiklus S, Sellers E, Krystal JH, Rothman DL, Mason GF. Preliminary evidence of reduced occipital GABA concentrations in puerperal women: a 1H-MRS study. Psychopharmacology (Berl) 2006;186:425–433. doi: 10.1007/s00213-006-0313-7. [DOI] [PubMed] [Google Scholar]
- Epperson CN, Haga K, Mason GF, Sellers E, Gueorguieva R, Zhang W, Weiss E, Rothman DL, Krystal JH. Cortical gamma-aminobutyric acid levels across the menstrual cycle in healthy women and those with premenstrual dysphoric disorder: a proton magnetic resonance spectroscopy study. Arch Gen Psychiatry. 2002;59:851–858. doi: 10.1001/archpsyc.59.9.851. [DOI] [PubMed] [Google Scholar]
- Epperson CN, O’Malley S, Czarkowski KA, Gueorguieva R, Jatlow P, Sanacora G, Rothman DL, Krystal JH, Mason GF. Sex, GABA, and nicotine: the impact of smoking on cortical GABA levels across the menstrual cycle as measured with proton magnetic resonance spectroscopy. Biol Psychiatry. 2005;57:44–48. doi: 10.1016/j.biopsych.2004.09.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Erickson SL, Lewis DA. Postnatal development of parvalbumin- and GABA transporter-immunoreactive axon terminals in monkey prefrontal cortex. J Comp Neurol. 2002;448:186–202. doi: 10.1002/cne.10249. [DOI] [PubMed] [Google Scholar]
- Errante LD, Williamson A, Spencer DD, Petroff OA. Gabapentin and vigabatrin increase GABA in the human neocortical slice. Epilepsy Res. 2002;49:203–210. doi: 10.1016/s0920-1211(02)00034-7. [DOI] [PubMed] [Google Scholar]
- Evans CJ, McGonigle DJ, Edden RA. Diurnal stability of gamma-aminobutyric acid concentration in visual and sensorimotor cortex. J Magn Reson Imaging. 2010;31:204–209. doi: 10.1002/jmri.21996. [DOI] [PubMed] [Google Scholar]
- Fabio MC, Nizhnikov ME, Spear NE, Pautassi RM. Binge ethanol intoxication heightens subsequent ethanol intake in adolescent, but not adult, rats. Dev Psychobiol. 2013 doi: 10.1002/dev.21101. [DOI] [PubMed] [Google Scholar]
- Faingold CL, N’Gouemo P, Riaz A. Ethanol and neurotransmitter interactions--from molecular to integrative effects. Prog Neurobiol. 1998;55:509–535. doi: 10.1016/s0301-0082(98)00027-6. [DOI] [PubMed] [Google Scholar]
- Farrant M, Kaila K. The cellular, molecular and ionic basis of GABA(A) receptor signalling. Prog Brain Res. 2007;160:59–87. doi: 10.1016/S0079-6123(06)60005-8. [DOI] [PubMed] [Google Scholar]
- Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABA(A) receptors. Nat Rev Neurosci. 2005;6:215–229. doi: 10.1038/nrn1625. [DOI] [PubMed] [Google Scholar]
- Fehr C, Sander T, Tadic A, Lenzen KP, Anghelescu I, Klawe C, Dahmen N, Schmidt LG, Szegedi A. Confirmation of association of the GABRA2 gene with alcohol dependence by subtype-specific analysis. Psychiatr Genet. 2006;16:9–17. doi: 10.1097/01.ypg.0000185027.89816.d9. [DOI] [PubMed] [Google Scholar]
- Ferrara SD, Zancaner S, Giorgetti R. Low blood alcohol concentrations and driving impairment. A review of experimental studies and international legislation. Int J Legal Med. 1994;106:169–177. doi: 10.1007/BF01371332. [DOI] [PubMed] [Google Scholar]
- Fillmore MT, Marczinski CA, Bowman AM. Acute tolerance to alcohol effects on inhibitory and activational mechanisms of behavioral control. J Stud Alcohol. 2005;66:663–672. doi: 10.15288/jsa.2005.66.663. [DOI] [PubMed] [Google Scholar]
- Fleming RL, Acheson SK, Moore SD, Wilson WA, Swartzwelder HS. In the rat, chronic intermittent ethanol exposure during adolescence alters the ethanol sensitivity of tonic inhibition in adulthood. Alcohol Clin Exp Res. 2012;36:279–285. doi: 10.1111/j.1530-0277.2011.01615.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fleming RL, Wilson WA, Swartzwelder HS. Magnitude and ethanol sensitivity of tonic GABAA receptor-mediated inhibition in dentate gyrus changes from adolescence to adulthood. J Neurophysiol. 2007;97:3806–3811. doi: 10.1152/jn.00101.2007. [DOI] [PubMed] [Google Scholar]
- Fritschy JM, Paysan J, Enna A, Mohler H. Switch in the expression of rat GABAA-receptor subtypes during postnatal development: an immunohistochemical study. J Neurosci. 1994;14:5302–5324. doi: 10.1523/JNEUROSCI.14-09-05302.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fung SJ, Webster MJ, Sivagnanasundaram S, Duncan C, Elashoff M, Weickert CS. Expression of interneuron markers in the dorsolateral prefrontal cortex of the developing human and in schizophrenia. Am J Psychiatry. 2010;167:1479–1488. doi: 10.1176/appi.ajp.2010.09060784. [DOI] [PubMed] [Google Scholar]
- Gambarana C, Beattie CE, Rodriguez ZR, Siegel RE. Region-specific expression of messenger RNAs encoding GABAA receptor subunits in the developing rat brain. Neuroscience. 1991;45:423–432. doi: 10.1016/0306-4522(91)90238-j. [DOI] [PubMed] [Google Scholar]
- Gann H, Feige B, Hohagen F, van Calker D, Geiss D, Dieter R. Sleep and the cholinergic rapid eye movement sleep induction test in patients with primary alcohol dependence. Biol Psychiatry. 2001;50:383–390. doi: 10.1016/s0006-3223(01)01172-6. [DOI] [PubMed] [Google Scholar]
- Gao F, Edden RA, Li M, Puts NA, Wang G, Liu C, Zhao B, Wang H, Bai X, Zhao C, Wang X, Barker PB. Edited magnetic resonance spectroscopy detects an age-related decline in brain GABA levels. Neuroimage. 2013 doi: 10.1016/j.neuroimage.2013.04.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giedd JN. The teen brain: insights from neuroimaging. J Adolesc Health. 2008;42:335–343. doi: 10.1016/j.jadohealth.2008.01.007. [DOI] [PubMed] [Google Scholar]
- Giedd JN, Blumenthal J, Jeffries NO, Castellanos FX, Liu H, Zijdenbos A, Paus T, Evans AC, Rapoport JL. Brain development during childhood and adolescence: a longitudinal MRI study. Nat Neurosci. 1999a;2:861–863. doi: 10.1038/13158. [DOI] [PubMed] [Google Scholar]
- Giedd JN, Blumenthal J, Jeffries NO, Rajapakse JC, Vaituzis AC, Liu H, Berry YC, Tobin M, Nelson J, Castellanos FX. Development of the human corpus callosum during childhood and adolescence: a longitudinal MRI study. Prog Neuropsychopharmacol Biol Psychiatry. 1999b;23:571–588. doi: 10.1016/s0278-5846(99)00017-2. [DOI] [PubMed] [Google Scholar]
- Giedd JN, Snell JW, Lange N, Rajapakse JC, Casey BJ, Kozuch PL, Vaituzis AC, Vauss YC, Hamburger SD, Kaysen D, Rapoport JL. Quantitative magnetic resonance imaging of human brain development: ages 4–18. Cereb Cortex. 1996a;6:551–560. doi: 10.1093/cercor/6.4.551. [DOI] [PubMed] [Google Scholar]
- Giedd JN, Vaituzis AC, Hamburger SD, Lange N, Rajapakse JC, Kaysen D, Vauss YC, Rapoport JL. Quantitative MRI of the temporal lobe, amygdala, and hippocampus in normal human development: ages 4–18 years. J Comp Neurol. 1996b;366:223–230. doi: 10.1002/(SICI)1096-9861(19960304)366:2<223::AID-CNE3>3.0.CO;2-7. [DOI] [PubMed] [Google Scholar]
- Goddard AW, Mason GF, Almai A, Rothman DL, Behar KL, Petroff OA, Charney DS, Krystal JH. Reductions in occipital cortex GABA levels in panic disorder detected with 1h-magnetic resonance spectroscopy. Arch Gen Psychiatry. 2001;58:556–561. doi: 10.1001/archpsyc.58.6.556. [DOI] [PubMed] [Google Scholar]
- Goddard AW, Mason GF, Appel M, Rothman DL, Gueorguieva R, Behar KL, Krystal JH. Impaired GABA neuronal response to acute benzodiazepine administration in panic disorder. Am J Psychiatry. 2004;161:2186–2193. doi: 10.1176/appi.ajp.161.12.2186. [DOI] [PubMed] [Google Scholar]
- Gogtay N, Giedd JN, Lusk L, Hayashi KM, Greenstein D, Vaituzis AC, Nugent TF, 3rd, Herman DH, Clasen LS, Toga AW, Rapoport JL, Thompson PM. Dynamic mapping of human cortical development during childhood through early adulthood. Proc Natl Acad Sci U S A. 2004;101:8174–8179. doi: 10.1073/pnas.0402680101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gogtay N, Nugent TF, 3rd, Herman DH, Ordonez A, Greenstein D, Hayashi KM, Clasen L, Toga AW, Giedd JN, Rapoport JL, Thompson PM. Dynamic mapping of normal human hippocampal development. Hippocampus. 2006;16:664–672. doi: 10.1002/hipo.20193. [DOI] [PubMed] [Google Scholar]
- Gogtay N, Thompson PM. Mapping gray matter development: implications for typical development and vulnerability to psychopathology. Brain Cogn. 2010;72:6–15. doi: 10.1016/j.bandc.2009.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gomez R, Behar KL, Watzl J, Weinzimer SA, Gulanski B, Sanacora G, Koretski J, Guidone E, Jiang L, Petrakis IL, Pittman B, Krystal JH, Mason GF. Intravenous ethanol infusion decreases human cortical gamma-aminobutyric acid and N-acetylaspartate as measured with proton magnetic resonance spectroscopy at 4 tesla. Biol Psychiatry. 2012;71:239–246. doi: 10.1016/j.biopsych.2011.06.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goto N, Yoshimura R, Moriya J, Kakeda S, Hayashi K, Ueda N, Ikenouchi-Sugita A, Umene-Nakano W, Oonari N, Korogi Y, Nakamura J. Critical examination of a correlation between brain gamma-aminobutyric acid (GABA) concentrations and a personality trait of extroversion in healthy volunteers as measured by a 3 Tesla proton magnetic resonance spectroscopy study. Psychiatry Res. 2010;182:53–57. doi: 10.1016/j.pscychresns.2009.11.002. [DOI] [PubMed] [Google Scholar]
- Goudriaan AE, Grekin ER, Sher KJ. Decision making and binge drinking: a longitudinal study. Alcohol Clin Exp Res. 2007;31:928–938. doi: 10.1111/j.1530-0277.2007.00378.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grant BF, Dawson DA. Age at onset of alcohol use and its association with DSM-IV alcohol abuse and dependence: results from the National Longitudinal Alcohol Epidemiologic Survey. J Subst Abuse. 1997;9:103–110. doi: 10.1016/s0899-3289(97)90009-2. [DOI] [PubMed] [Google Scholar]
- Grobin AC, Matthews DB, Devaud LL, Morrow AL. The role of GABA(A) receptors in the acute and chronic effects of ethanol. Psychopharmacology (Berl) 1998;139:2–19. doi: 10.1007/s002130050685. [DOI] [PubMed] [Google Scholar]
- Grusser SM, Wrase J, Klein S, Hermann D, Smolka MN, Ruf M, Weber-Fahr W, Flor H, Mann K, Braus DF, Heinz A. Cue-induced activation of the striatum and medial prefrontal cortex is associated with subsequent relapse in abstinent alcoholics. Psychopharmacology (Berl) 2004;175:296–302. doi: 10.1007/s00213-004-1828-4. [DOI] [PubMed] [Google Scholar]
- Guerri C, Pascual M. Mechanisms involved in the neurotoxic, cognitive, and neurobehavioral effects of alcohol consumption during adolescence. Alcohol. 2010;44:15–26. doi: 10.1016/j.alcohol.2009.10.003. [DOI] [PubMed] [Google Scholar]
- Guo Y, Kaplan IV, Cooper NG, Mower GD. Expression of two forms of glutamic acid decarboxylase (GAD67 and GAD65) during postnatal development of the cat visual cortex. Brain Res Dev Brain Res. 1997;103:127–141. doi: 10.1016/s0165-3806(97)81789-0. [DOI] [PubMed] [Google Scholar]
- Gutierrez A, Khan ZU, Miralles CP, Mehta AK, Ruano D, Araujo F, Vitorica J, De Blas AL. GABAA receptor subunit expression changes in the rat cerebellum and cerebral cortex during aging. Brain Res Mol Brain Res. 1997;45:59–70. doi: 10.1016/s0169-328x(96)00237-9. [DOI] [PubMed] [Google Scholar]
- Ham BJ, Sung Y, Kim N, Kim SJ, Kim JE, Kim DJ, Lee JY, Kim JH, Yoon SJ, Lyoo IK. Decreased GABA levels in anterior cingulate and basal ganglia in medicated subjects with panic disorder: a proton magnetic resonance spectroscopy (1H-MRS) study. Prog Neuropsychopharmacol Biol Psychiatry. 2007;31:403–411. doi: 10.1016/j.pnpbp.2006.10.011. [DOI] [PubMed] [Google Scholar]
- Harris GJ, Jaffin SK, Hodge SM, Kennedy D, Caviness VS, Marinkovic K, Papadimitriou GM, Makris N, Oscar-Berman M. Frontal white matter and cingulum diffusion tensor imaging deficits in alcoholism. Alcohol Clin Exp Res. 2008a;32:1001–1013. doi: 10.1111/j.1530-0277.2008.00661.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harris RA, Trudell JR, Mihic SJ. Ethanol’s molecular targets. Sci Signal. 2008b;1:re7. doi: 10.1126/scisignal.128re7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hasler G, van der Veen JW, Geraci M, Shen J, Pine D, Drevets WC. Prefrontal cortical gamma-aminobutyric Acid levels in panic disorder determined by proton magnetic resonance spectroscopy. Biol Psychiatry. 2009;65:273–275. doi: 10.1016/j.biopsych.2008.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heidbreder C, De Witte P. Ethanol differentially affects extracellular monoamines and GABA in the nucleus accumbens. Pharmacol Biochem Behav. 1993;46:477–481. doi: 10.1016/0091-3057(93)90383-5. [DOI] [PubMed] [Google Scholar]
- Heinz A, Wrase J, Kahnt T, Beck A, Bromand Z, Grusser SM, Kienast T, Smolka MN, Flor H, Mann K. Brain activation elicited by affectively positive stimuli is associated with a lower risk of relapse in detoxified alcoholic subjects. Alcohol Clin Exp Res. 2007;31:1138–1147. doi: 10.1111/j.1530-0277.2007.00406.x. [DOI] [PubMed] [Google Scholar]
- Henschel O, Gipson KE, Bordey A. GABAA receptors, anesthetics and anticonvulsants in brain development. CNS Neurol Disord Drug Targets. 2008;7:211–224. doi: 10.2174/187152708784083812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herting MM, Maxwell EC, Irvine C, Nagel BJ. The impact of sex, puberty, and hormones on white matter microstructure in adolescents. Cereb Cortex. 2012;22:1979–1992. doi: 10.1093/cercor/bhr246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hetherington HP, Newcomer BR, Pan JW. Measurements of human cerebral GABA at 4.1 T using numerically optimized editing pulses. Magnetic Resonance in Medicine. 1998;39:6–10. doi: 10.1002/mrm.1910390103. [DOI] [PubMed] [Google Scholar]
- Hill SY. Trajectories of alcohol use and electrophysiological and morphological indices of brain development: distinguishing causes from consequences. Ann N Y Acad Sci. 2004;1021:245–259. doi: 10.1196/annals.1308.029. [DOI] [PubMed] [Google Scholar]
- Hill SY, Muddasani S, Prasad K, Nutche J, Steinhauer SR, Scanlon J, McDermott M, Keshavan M. Cerebellar volume in offspring from multiplex alcohol dependence families. Biol Psychiatry. 2007;61:41–47. doi: 10.1016/j.biopsych.2006.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hill SY, Shen S, Lowers L, Locke J. Factors predicting the onset of adolescent drinking in families at high risk for developing alcoholism. Biol Psychiatry. 2000;48:265–275. doi: 10.1016/s0006-3223(00)00841-6. [DOI] [PubMed] [Google Scholar]
- Hingson RW, Heeren T, Winter MR. Age at drinking onset and alcohol dependence: age at onset, duration, and severity. Arch Pediatr Adolesc Med. 2006a;160:739–746. doi: 10.1001/archpedi.160.7.739. [DOI] [PubMed] [Google Scholar]
- Hingson RW, Heeren T, Winter MR. Age at drinking onset and alcohol dependence: age at onset, duration, and severity. Arch Ped Adol Med. 2006b;160:739–746. doi: 10.1001/archpedi.160.7.739. [DOI] [PubMed] [Google Scholar]
- Hollenstein T, Lougheed JP. Beyond storm and stress: Typicality, transactions, timing, and temperament to account for adolescent change. Am Psychol. 2013;68:444–454. doi: 10.1037/a0033586. [DOI] [PubMed] [Google Scholar]
- Hollstedt C, Olsson O, Rydberg U. Effects of ethanol on the developing rat. II. Coordination as measured by the tilting-plane test. Med Biol. 1980;58:164–168. [PubMed] [Google Scholar]
- Hu X, Oroszi G, Chun J, Smith TL, Goldman D, Schuckit MA. An expanded evaluation of the relationship of four alleles to the level of response to alcohol and the alcoholism risk. Alcohol Clin Exp Res. 2005;29:8–16. doi: 10.1097/01.alc.0000150008.68473.62. [DOI] [PubMed] [Google Scholar]
- Hulin MW, Amato RJ, Winsauer PJ. GABA(A) receptor modulation during adolescence alters adult ethanol intake and preference in rats. Alcohol Clin Exp Res. 2012;36:223–233. doi: 10.1111/j.1530-0277.2011.01622.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hwang BH, Kunkler PE, Lumeng L. Quantitative Autoradiography on [(35)S]TBPS Binding Sites of Gamma- Aminobutyric Acid(A) Receptors in Discrete Brain Regions of High- Alcohol-Drinking and Low-Alcohol- Drinking Rats Selectively Bred forHigh- and Low-Alcohol Preference. J Biomed Sci. 1997;4:308–314. doi: 10.1007/BF02258355. [DOI] [PubMed] [Google Scholar]
- Jensen JE, de Frederick BB, Renshaw PF. Grey and white matter GABA level differences in the human brain using two-dimensional, J-resolved spectroscopic imaging. NMR Biomed. 2005;18:570–576. doi: 10.1002/nbm.994. [DOI] [PubMed] [Google Scholar]
- Jiang M, Oliva AA, Jr, Lam T, Swann JW. GABAergic neurons that pioneer hippocampal area CA1 of the mouse: morphologic features and multiple fates. J Comp Neurol. 2001;439:176–192. doi: 10.1002/cne.1341. [DOI] [PubMed] [Google Scholar]
- Jocham G, Hunt LT, Near J, Behrens TE. A mechanism for value-guided choice based on the excitation-inhibition balance in prefrontal cortex. Nat Neurosci. 2012;15:960–961. doi: 10.1038/nn.3140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnston LD, O’Malley PM, Bachman JG, Schulenberg JE. Monitoring the Future national survey results on adolescent drug use: Overview of key findings, 2010. Ann Arbor: Institute for Social Research, The University of Michigan; 2011. [Google Scholar]
- Johnston MV, Coyle JT. Development of central neurotransmitter systems. Ciba Found Symp. 1981;86:251–270. doi: 10.1002/9780470720684.ch12. [DOI] [PubMed] [Google Scholar]
- Kakeda S, Korogi Y, Moriya J, Ohnari N, Sato T, Ueno S, Yanagihara N, Harada M, Matsuda T. Influence of work shift on glutamic acid and gamma-aminobutyric acid (GABA): evaluation with proton magnetic resonance spectroscopy at 3T. Psychiatry Res. 2011;192:55–59. doi: 10.1016/j.pscychresns.2010.10.011. [DOI] [PubMed] [Google Scholar]
- Kawaguchi Y, Kondo S. Parvalbumin, somatostatin and cholecystokinin as chemical markers for specific GABAergic interneuron types in the rat frontal cortex. J Neurocytol. 2002;31:277–287. doi: 10.1023/a:1024126110356. [DOI] [PubMed] [Google Scholar]
- Kegeles LS, Mao X, Stanford AD, Girgis R, Ojeil N, Xu X, Gil R, Slifstein M, Abi-Dargham A, Lisanby SH, Shungu DC. Elevated prefrontal cortex gamma-aminobutyric acid and glutamate-glutamine levels in schizophrenia measured in vivo with proton magnetic resonance spectroscopy. Arch Gen Psychiatry. 2012;69:449–459. doi: 10.1001/archgenpsychiatry.2011.1519. [DOI] [PubMed] [Google Scholar]
- Keltner JR, Wald LL, Frederick BB, Renshaw PF. In vivo detection of GABA in human brain using a localized double-quantum filter technique. Magn Reson Med. 1997;37:366–371. doi: 10.1002/mrm.1910370312. [DOI] [PubMed] [Google Scholar]
- Kemppainen H, Raivio N, Nurmi H, Kiianmaa K. GABA and glutamate overflow in the VTA and ventral pallidum of alcohol-preferring AA and alcohol-avoiding ANA rats after ethanol. Alcohol Alcohol. 2010;45:111–118. doi: 10.1093/alcalc/agp086. [DOI] [PubMed] [Google Scholar]
- Kilb W. Development of the GABAergic system from birth to adolescence. Neuroscientist. 2012;18:613–630. doi: 10.1177/1073858411422114. [DOI] [PubMed] [Google Scholar]
- Kim HJ, Kim JE, Cho G, Song IC, Bae S, Hong SJ, Yoon SJ, Lyoo IK, Kim TS. Associations between anterior cingulate cortex glutamate and gamma-aminobutyric acid concentrations and the harm avoidance temperament. Neurosci Lett. 2009;464:103–107. doi: 10.1016/j.neulet.2009.07.087. [DOI] [PubMed] [Google Scholar]
- Klenberg L, Korkman M, Lahti-Nuuttila P. Differential development of attention and executive functions in 3- to 12-year-old Finnish children. Dev Neuropsychol. 2001;20:407–428. doi: 10.1207/S15326942DN2001_6. [DOI] [PubMed] [Google Scholar]
- Korpi ER, Debus F, Linden AM, Malecot C, Leppa E, Vekovischeva O, Rabe H, Bohme I, Aller MI, Wisden W, Luddens H. Does ethanol act preferentially via selected brain GABAA receptor subtypes? the current evidence is ambiguous. Alcohol. 2007;41:163–176. doi: 10.1016/j.alcohol.2007.03.007. [DOI] [PubMed] [Google Scholar]
- Kristt DA, Waldman JV. Late postnatal changes in rat somatosensory cortex. Temporal and spatial relationships of GABA-T and AChE histochemical reactivity. Anat Embryol (Berl) 1986;174:115–122. doi: 10.1007/BF00318343. [DOI] [PubMed] [Google Scholar]
- Krystal JH, D’Souza DC, Gallinat J, Driesen N, Abi-Dargham A, Petrakis I, Heinz A, Pearlson G. The vulnerability to alcohol and substance abuse in individuals diagnosed with schizophrenia. Neurotox Res. 2006;10:235–252. doi: 10.1007/BF03033360. [DOI] [PubMed] [Google Scholar]
- Krystal JH, Sanacora G, Blumberg H, Anand A, Charney DS, Marek G, Epperson CN, Goddard A, Mason GF. Glutamate and GABA systems as targets for novel antidepressant and mood-stabilizing treatments. Mol Psychiatry. 2002;7(Suppl 1):S71–80. doi: 10.1038/sj.mp.4001021. [DOI] [PubMed] [Google Scholar]
- Kumar S, Porcu P, Werner DF, Matthews DB, Diaz-Granados JL, Helfand RS, Morrow AL. The role of GABA(A) receptors in the acute and chronic effects of ethanol: a decade of progress. Psychopharmacology (Berl) 2009;205:529–564. doi: 10.1007/s00213-009-1562-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Land C, Spear NE. Ethanol impairs memory of a simple discrimination in adolescent rats at doses that leave adult memory unaffected. Neurobiol Learn Mem. 2004a;81:75–81. doi: 10.1016/j.nlm.2003.08.005. [DOI] [PubMed] [Google Scholar]
- Land C, Spear NE. Fear conditioning is impaired in adult rats by ethanol doses that do not affect periadolescents. Int J Dev Neurosci. 2004b;22:355–362. doi: 10.1016/j.ijdevneu.2004.04.008. [DOI] [PubMed] [Google Scholar]
- Lappalainen J, Krupitsky E, Remizov M, Pchelina S, Taraskina A, Zvartau E, Somberg LK, Covault J, Kranzler HR, Krystal JH, Gelernter J. Association between alcoholism and gamma-amino butyric acid alpha2 receptor subtype in a Russian population. Alcohol Clin Exp Res. 2005;29:493–498. doi: 10.1097/01.alc.0000158938.97464.90. [DOI] [PubMed] [Google Scholar]
- Laurie DJ, Wisden W, Seeburg PH. The distribution of thirteen GABAA receptor subunit mRNAs in the rat brain. III. Embryonic and postnatal development. J Neurosci. 1992;12:4151–4172. doi: 10.1523/JNEUROSCI.12-11-04151.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leggio L, Kenna GA, Swift RM. New developments for the pharmacological treatment of alcohol withdrawal syndrome. A focus on non-benzodiazepine GABAergic medications. Prog Neuropsychopharmacol Biol Psychiatry. 2008;32:1106–1117. doi: 10.1016/j.pnpbp.2007.09.021. [DOI] [PubMed] [Google Scholar]
- Li Q, Wilson WA, Swartzwelder HS. Differential effect of ethanol on NMDA EPSCs in pyramidal cells in the posterior cingulate cortex of juvenile and adult rats. J Neurophysiol. 2002;87:705–711. doi: 10.1152/jn.00433.2001. [DOI] [PubMed] [Google Scholar]
- Li Q, Wilson WA, Swartzwelder HS. Developmental differences in the sensitivity of hippocampal GABAA receptor-mediated IPSCS to ethanol. Alcohol Clin Exp Res. 2003;27:2017–2022. doi: 10.1097/01.ALC.0000108390.62394.71. [DOI] [PubMed] [Google Scholar]
- Li Q, Wilson WA, Swartzwelder HS. Developmental differences in the sensitivity of spontaneous and miniature IPSCs to ethanol. Alcohol Clin Exp Res. 2006;30:119–126. doi: 10.1111/j.1530-0277.2006.00006.x. [DOI] [PubMed] [Google Scholar]
- Licata SC, Jensen JE, Penetar DM, Prescot AP, Lukas SE, Renshaw PF. A therapeutic dose of zolpidem reduces thalamic GABA in healthy volunteers: a proton MRS study at 4 T. Psychopharmacology (Berl) 2009;203:819–829. doi: 10.1007/s00213-008-1431-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lidow MS, Goldman-Rakic PS, Rakic P. Synchronized overproduction of neurotransmitter receptors in diverse regions of the primate cerebral cortex. Proc Natl Acad Sci U S A. 1991;88:10218–10221. doi: 10.1073/pnas.88.22.10218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liljequist S, Engel J. Effects of GABAergic agonists and antagonists on various ethanol-induced behavioral changes. Psychopharmacology (Berl) 1982;78:71–75. doi: 10.1007/BF00470592. [DOI] [PubMed] [Google Scholar]
- Lingford-Hughes AR, Acton PD, Gacinovic S, Boddington SJ, Costa DC, Pilowsky LS, Ell PJ, Marshall EJ, Kerwin RW. Levels of gamma-aminobutyric acid-benzodiazepine receptors in abstinent, alcohol-dependent women: preliminary findings from an 123I-iomazenil single photon emission tomography study. Alcohol Clin Exp Res. 2000;24:1449–1455. [PubMed] [Google Scholar]
- Lingford-Hughes AR, Acton PD, Gacinovic S, Suckling J, Busatto GF, Boddington SJ, Bullmore E, Woodruff PW, Costa DC, Pilowsky LS, Ell PJ, Marshall EJ, Kerwin RW. Reduced levels of GABA-benzodiazepine receptor in alcohol dependency in the absence of grey matter atrophy. Br J Psychiatry. 1998;173:116–122. doi: 10.1192/bjp.173.2.116. [DOI] [PubMed] [Google Scholar]
- Lingford-Hughes AR, Wilson SJ, Cunningham VJ, Feeney A, Stevenson B, Brooks DJ, Nutt DJ. GABA-benzodiazepine receptor function in alcohol dependence: a combined 11C-flumazenil PET and pharmacodynamic study. Psychopharmacology (Berl) 2005;180:595–606. doi: 10.1007/s00213-005-2271-x. [DOI] [PubMed] [Google Scholar]
- Little PJ, Kuhn CM, Wilson WA, Swartzwelder HS. Differential effects of ethanol in adolescent and adult rats. Alcohol Clin Exp Res. 1996;20:1346–1351. doi: 10.1111/j.1530-0277.1996.tb01133.x. [DOI] [PubMed] [Google Scholar]
- Lobo IA, Harris RA. GABA(A) receptors and alcohol. Pharmacol Biochem Behav. 2008;90:90–94. doi: 10.1016/j.pbb.2008.03.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Long Z, Medlock C, Dzemidzic M, Shin YW, Goddard AW, Dydak U. Decreased GABA levels in anterior cingulate cortex/medial prefrontal cortex in panic disorder. Prog Neuropsychopharmacol Biol Psychiatry. 2013;44C:131–135. doi: 10.1016/j.pnpbp.2013.01.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lovinger DM, Homanics GE. Tonic for what ails us? high-affinity GABAA receptors and alcohol. Alcohol. 2007;41:139–143. doi: 10.1016/j.alcohol.2007.03.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luna B, Sweeney JA. The emergence of collaborative brain function: FMRI studies of the development of response inhibition. Ann N Y Acad Sci. 2004;1021:296–309. doi: 10.1196/annals.1308.035. [DOI] [PubMed] [Google Scholar]
- Marinkovic K, Oscar-Berman M, Urban T, O’Reilly CE, Howard JA, Sawyer K, Harris GJ. Alcoholism and dampened temporal limbic activation to emotional faces. Alcohol Clin Exp Res. 2009;33:1880–1892. doi: 10.1111/j.1530-0277.2009.01026.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Markwiese BJ, Acheson SK, Levin ED, Wilson WA, Swartzwelder HS. Differential effects of ethanol on memory in adolescent and adult rats. Alcohol Clin Exp Res. 1998;22:416–421. [PubMed] [Google Scholar]
- Martin TL, Solbeck PA, Mayers DJ, Langille RM, Buczek Y, Pelletier MR. A review of alcohol-impaired driving: the role of blood alcohol concentration and complexity of the driving task. J Forensic Sci. 2013;58:1238–1250. doi: 10.1111/1556-4029.12227. [DOI] [PubMed] [Google Scholar]
- Mason GF, Petrakis IL, de Graaf RA, Gueorguieva R, Guidone E, Coric V, Epperson CN, Rothman DL, Krystal JH. Cortical gamma-aminobutyric acid levels and the recovery from ethanol dependence: preliminary evidence of modification by cigarette smoking. Biol Psychiatry. 2006;59:85–93. doi: 10.1016/j.biopsych.2005.06.009. [DOI] [PubMed] [Google Scholar]
- Masten AS, Faden VB, Zucker RA, Spear LP. Underage drinking: a developmental framework. Pediatrics. 2008;121(Suppl 4):S235–251. doi: 10.1542/peds.2007-2243A. [DOI] [PubMed] [Google Scholar]
- Matthews AG, Hoffman EK, Zezza N, Stiffler S, Hill SY. The role of the GABRA2 polymorphism in multiplex alcohol dependence families with minimal comorbidity: within-family association and linkage analyses. J Stud Alcohol Drugs. 2007;68:625–633. doi: 10.15288/jsad.2007.68.625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matthews DB. Adolescence and alcohol: recent advances in understanding the impact of alcohol use during a critical developmental window. Alcohol. 2010;44:1–2. doi: 10.1016/j.alcohol.2009.10.018. [DOI] [PubMed] [Google Scholar]
- Matthews DB, Tinsley KL, Diaz-Granados JL, Tokunaga S, Silvers JM. Chronic intermittent exposure to ethanol during adolescence produces tolerance to the hypnotic effects of ethanol in male rats: a dose-dependent analysis. Alcohol. 2008;42:617–621. doi: 10.1016/j.alcohol.2008.09.001. [DOI] [PubMed] [Google Scholar]
- McCormick DA. GABA as an inhibitory neurotransmitter in human cerebral cortex. J Neurophysiol. 1989;62:1018–1027. doi: 10.1152/jn.1989.62.5.1018. [DOI] [PubMed] [Google Scholar]
- McDonald JW, Johnston MV. Physiological and pathophysiological roles of excitatory amino acids during central nervous system development. Brain Res Brain Res Rev. 1990;15:41–70. doi: 10.1016/0165-0173(90)90011-c. [DOI] [PubMed] [Google Scholar]
- McQueeny T, Schweinsburg BC, Schweinsburg AD, Jacobus J, Bava S, Frank LR, Tapert SF. Altered white matter integrity in adolescent binge drinkers. Alcohol Clin Exp Res. 2009;33:1278–1285. doi: 10.1111/j.1530-0277.2009.00953.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mekle R, Mlynarik V, Gambarota G, Hergt M, Krueger G, Gruetter R. MR spectroscopy of the human brain with enhanced signal intensity at ultrashort echo times on a clinical platform at 3T and 7T. Magn Reson Med. 2009;61:1279–1285. doi: 10.1002/mrm.21961. [DOI] [PubMed] [Google Scholar]
- Mescher M, Merkle H, Kirsch J, Garwood M, Gruetter R. Simultaneous in vivo spectral editing and water suppression. NMR Biomed. 1998;11:266–272. doi: 10.1002/(sici)1099-1492(199810)11:6<266::aid-nbm530>3.0.co;2-j. [DOI] [PubMed] [Google Scholar]
- Meyerhoff DJ, Blumenfeld R, Truran D, Lindgren J, Flenniken D, Cardenas V, Chao LL, Rothlind J, Studholme C, Weiner MW. Effects of heavy drinking, binge drinking, and family history of alcoholism on regional brain metabolites. Alcohol Clin Exp Res. 2004;28:650–661. doi: 10.1097/01.ALC.0000121805.12350.CA. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meyerhoff DJ, Durazzo TC, Ende G. Chronic alcohol consumption, abstinence and relapse: brain proton magnetic resonance spectroscopy studies in animals and humans. Curr Top Behav Neurosci. 2013;13:511–540. doi: 10.1007/7854_2011_131. [DOI] [PubMed] [Google Scholar]
- Minelli A, Barbaresi P, Conti F. Postnatal development of high-affinity plasma membrane GABA transporters GAT-2 and GAT-3 in the rat cerebral cortex. Brain Res Dev Brain Res. 2003;142:7–18. doi: 10.1016/s0165-3806(03)00007-5. [DOI] [PubMed] [Google Scholar]
- Minier F, Sigel E. Positioning of the alpha-subunit isoforms confers a functional signature to gamma-aminobutyric acid type A receptors. Proc Natl Acad Sci U S A. 2004;101:7769–7774. doi: 10.1073/pnas.0400220101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Modig F, Fransson PA, Magnusson M, Patel M. Blood alcohol concentration at 0.06 and 0.10% causes a complex multifaceted deterioration of body movement control. Alcohol. 2012;46:75–88. doi: 10.1016/j.alcohol.2011.06.001. [DOI] [PubMed] [Google Scholar]
- Mokdad AH, Marks JS, Stroup DF, Gerberding JL. Actual causes of death in the United States, 2000. JAMA. 2004;291:1238–1245. doi: 10.1001/jama.291.10.1238. [DOI] [PubMed] [Google Scholar]
- Mon A, Durazzo TC, Meyerhoff DJ. Glutamate, GABA, and other cortical metabolite concentrations during early abstinence from alcohol and their associations with neurocognitive changes. Drug Alcohol Depend. 2012;125:27–36. doi: 10.1016/j.drugalcdep.2012.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morgan PT, Pace-Schott EF, Mason GF, Forselius E, Fasula M, Valentine GW, Sanacora G. Cortical GABA levels in primary insomnia. Sleep. 2012;35:807–814. doi: 10.5665/sleep.1880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morris SA, Kelso ML, Liput DJ, Marshall SA, Nixon K. Similar withdrawal severity in adolescents and adults in a rat model of alcohol dependence. Alcohol. 2010;44:89–98. doi: 10.1016/j.alcohol.2009.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morrow AL, VanDoren MJ, Penland SN, Matthews DB. The role of GABAergic neuroactive steroids in ethanol action, tolerance and dependence. Brain Res Brain Res Rev. 2001;37:98–109. doi: 10.1016/s0165-0173(01)00127-8. [DOI] [PubMed] [Google Scholar]
- Moy SS, Duncan GE, Knapp DJ, Breese GR. Sensitivity to ethanol across development in rats: comparison to [3H]zolpidem binding. Alcohol Clin Exp Res. 1998;22:1485–1492. [PubMed] [Google Scholar]
- Mueller SG, Weber OM, Duc CO, Weber B, Meier D, Russ W, Boesiger P, Wieser HG. Effects of vigabatrin on brain GABA+/CR signals in patients with epilepsy monitored by 1H-NMR-spectroscopy: responder characteristics. Epilepsia. 2001;42:29–40. doi: 10.1046/j.1528-1157.2001.077889.x. [DOI] [PubMed] [Google Scholar]
- Mukherjee S, Simasko SM. Chronic alcohol treatment in rats alters sleep by fragmenting periods of vigilance cycling in the light period with extended wakenings. Behav Brain Res. 2009;198:113–124. doi: 10.1016/j.bbr.2008.10.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mullins PG, McGonigle DJ, O’Gorman RL, Puts NA, Vidyasagar R, Evans CJ, Edden RA. Current practice in the use of MEGA-PRESS spectroscopy for the detection of GABA. Neuroimage. 2012 doi: 10.1016/j.neuroimage.2012.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Myrick H, Anton RF, Li X, Henderson S, Drobes D, Voronin K, George MS. Differential brain activity in alcoholics and social drinkers to alcohol cues: relationship to craving. Neuropsychopharmacology. 2004;29:393–402. doi: 10.1038/sj.npp.1300295. [DOI] [PubMed] [Google Scholar]
- Nagel BJ, Medina KL, Yoshii J, Schweinsburg AD, Moadab I, Tapert SF. Age-related changes in prefrontal white matter volume across adolescence. Neuroreport. 2006;17:1427–1431. doi: 10.1097/01.wnr.0000233099.97784.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- NIAAA. NIAAA Newsletter. Vol. 3. Bethesda, MD: NIAAA; 2004. National Institute of Alcohol Abuse and Alcoholism council approves definition of binge drinking; p. 3. [Google Scholar]
- Nixon K, McClain JA. Adolescence as a critical window for developing an alcohol use disorder: current findings in neuroscience. Curr Opin Psychiatry. 2010;23:227–232. doi: 10.1097/YCO.0b013e32833864fe. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Norman AL, Pulido C, Squeglia LM, Spadoni AD, Paulus MP, Tapert SF. Neural activation during inhibition predicts initiation of substance use in adolescence. Drug Alcohol Depend. 2011;119:216–223. doi: 10.1016/j.drugalcdep.2011.06.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Northoff G, Walter M, Schulte RF, Beck J, Dydak U, Henning A, Boeker H, Grimm S, Boesiger P. GABA concentrations in the human anterior cingulate cortex predict negative BOLD responses in fMRI. Nat Neurosci. 2007;10:1515–1517. doi: 10.1038/nn2001. [DOI] [PubMed] [Google Scholar]
- Novier A, Skike CE, Chin VS, Diaz-Granados JL, Matthews DB. Low and moderate doses of acute ethanol do not impair spatial cognition but facilitate accelerating rotarod performance in adolescent and adult rats. Neurosci Lett. 2012;512:38–42. doi: 10.1016/j.neulet.2012.01.059. [DOI] [PubMed] [Google Scholar]
- Novotny EJ, Jr, Hyder F, Shevell M, Rothman DL. GABA changes with vigabatrin in the developing human brain. Epilepsia. 1999;40:462–466. doi: 10.1111/j.1528-1157.1999.tb00741.x. [DOI] [PubMed] [Google Scholar]
- O’Donnell P. Cortical interneurons, immune factors and oxidative stress as early targets for schizophrenia. Eur J Neurosci. 2012;35:1866–1870. doi: 10.1111/j.1460-9568.2012.08130.x. [DOI] [PubMed] [Google Scholar]
- Olsen RW, Sieghart W. International Union of Pharmacology. LXX. Subtypes of gamma-aminobutyric acid(A) receptors: classification on the basis of subunit composition, pharmacology, and function. Update. Pharmacol Rev. 2008;60:243–260. doi: 10.1124/pr.108.00505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Olsen RW, Sieghart W. GABA A receptors: subtypes provide diversity of function and pharmacology. Neuropharmacology. 2009;56:141–148. doi: 10.1016/j.neuropharm.2008.07.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Olsen RW, Spigelman I. GABAA Receptor Plasticity in Alcohol Withdrawal. 4. Bethesda: National Center for Biotechnology Information; 2012. [PubMed] [Google Scholar]
- Oscar-Berman M. Learning and memory deficits in detoxified alcoholics. NIDA Res Monogr. 1990;101:136–155. [PubMed] [Google Scholar]
- Oscar-Berman M. Neuropsychological vulnerabilities in chronic alcoholism. In: Noronha A, Eckardt M, Warren K, editors. Review of NIAAA’s Neuroscience and Behavioral Research Portfolio. Bethesda, MD: National Institutes of Health; 2000. pp. 437–472. Vol. NIAAA Research Monograph No. 34. [Google Scholar]
- Oscar-Berman M, Ellis RJ. Cognitive deficits related to memory impairments in alcoholism. Recent Dev Alcohol. 1987;5:59–80. doi: 10.1007/978-1-4899-1684-6_3. [DOI] [PubMed] [Google Scholar]
- Oscar-Berman M, Marinkovic K. Alcoholism and the brain: an overview. Alcohol Res Health. 2003;27:125–133. [PMC free article] [PubMed] [Google Scholar]
- Oscar-Berman M, Marinkovic K. Alcohol: effects on neurobehavioral functions and the brain. Neuropsychol Rev. 2007;17:239–257. doi: 10.1007/s11065-007-9038-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parsons OA, Nixon SJ. Cognitive functioning in sober social drinkers: a review of the research since 1986. J Stud Alcohol. 1998;59:180–190. doi: 10.15288/jsa.1998.59.180. [DOI] [PubMed] [Google Scholar]
- Pasley BN, Inglis BA, Freeman RD. Analysis of oxygen metabolism implies a neural origin for the negative BOLD response in human visual cortex. Neuroimage. 2007;36:269–276. doi: 10.1016/j.neuroimage.2006.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paulus MP, Schuckit MA, Tapert SF, Tolentino NJ, Matthews SC, Smith TL, Trim RS, Hall SA, Simmons AN. High versus low level of response to alcohol: evidence of differential reactivity to emotional stimuli. Biol Psychiatry. 2012;72:848–855. doi: 10.1016/j.biopsych.2012.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paus T. Mapping brain maturation and cognitive development during adolescence. Trends Cogn Sci. 2005;9:60–68. doi: 10.1016/j.tics.2004.12.008. [DOI] [PubMed] [Google Scholar]
- Paus T, Keshavan M, Giedd JN. Why do many psychiatric disorders emerge during adolescence? Nat Rev Neurosci. 2008;9:947–957. doi: 10.1038/nrn2513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pautassi RM, Myers M, Spear LP, Molina JC, Spear NE. Adolescent but not adult rats exhibit ethanol-mediated appetitive second-order conditioning. Alcohol Clin Exp Res. 2008;32:2016–2027. doi: 10.1111/j.1530-0277.2008.00789.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Petroff OA, Behar KL, Mattson RH, Rothman DL. Human brain gamma-aminobutyric acid levels and seizure control following initiation of vigabatrin therapy. J Neurochem. 1996a;67:2399–2404. doi: 10.1046/j.1471-4159.1996.67062399.x. [DOI] [PubMed] [Google Scholar]
- Petroff OA, Hyder F, Rothman DL, Mattson RH. Topiramate rapidly raises brain GABA in epilepsy patients. Epilepsia. 2001;42:543–548. doi: 10.1046/j.1528-1157.2001.18800.x. [DOI] [PubMed] [Google Scholar]
- Petroff OA, Rothman DL, Behar KL, Hyder F, Mattson RH. Effects of valproate and other antiepileptic drugs on brain glutamate, glutamine, and GABA in patients with refractory complex partial seizures. Seizure. 1999;8:120–127. doi: 10.1053/seiz.1999.0267. [DOI] [PubMed] [Google Scholar]
- Petroff OA, Rothman DL, Behar KL, Mattson RH. Low brain GABA level is associated with poor seizure control. Ann Neurol. 1996b;40:908–911. doi: 10.1002/ana.410400613. [DOI] [PubMed] [Google Scholar]
- Pfefferbaum A, Mathalon DH, Sullivan EV, Rawles JM, Zipursky RB, Lim KO. A quantitative magnetic resonance imaging study of changes in brain morphology from infancy to late adulthood. Arch Neurol. 1994a;51:874–887. doi: 10.1001/archneur.1994.00540210046012. [DOI] [PubMed] [Google Scholar]
- Pfefferbaum A, Mathalon DH, Sullivan EV, Rawles JM, Zipursky RB, Lim KO. A quantitative magnetic resonance imaging study of changes in brain morphology from infancy to late adulthood. Arch Neurol. 1994b;51:874–887. doi: 10.1001/archneur.1994.00540210046012. [DOI] [PubMed] [Google Scholar]
- Pfefferbaum A, Sullivan EV, Mathalon DH, Lim KO. Frontal lobe volume loss observed with magnetic resonance imaging in older chronic alcoholics. Alcohol Clin Exp Res. 1997;21:521–529. doi: 10.1111/j.1530-0277.1997.tb03798.x. [DOI] [PubMed] [Google Scholar]
- Piepponen TP, Kiianmaa K, Ahtee L. Effects of ethanol on the accumbal output of dopamine, GABA and glutamate in alcohol-tolerant and alcohol-nontolerant rats. Pharmacol Biochem Behav. 2002;74:21–30. doi: 10.1016/s0091-3057(02)00937-1. [DOI] [PubMed] [Google Scholar]
- Pitts FN, Jr, Quick C. Brain succinate semialdehyde. II. Changes in the developing rat brain. J Neurochem. 1967;14:561–570. doi: 10.1111/j.1471-4159.1967.tb09556.x. [DOI] [PubMed] [Google Scholar]
- Plante DT, Jensen JE, Schoerning L, Winkelman JW. Reduced gamma-aminobutyric acid in occipital and anterior cingulate cortices in primary insomnia: a link to major depressive disorder? Neuropsychopharmacology. 2012a;37:1548–1557. doi: 10.1038/npp.2012.4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Plante DT, Jensen JE, Winkelman JW. The role of GABA in primary insomnia. Sleep. 2012b;35:741–742. doi: 10.5665/sleep.1854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollack MH, Jensen JE, Simon NM, Kaufman RE, Renshaw PF. High-field MRS study of GABA, glutamate and glutamine in social anxiety disorder: response to treatment with levetiracetam. Prog Neuropsychopharmacol Biol Psychiatry. 2008;32:739–743. doi: 10.1016/j.pnpbp.2007.11.023. [DOI] [PubMed] [Google Scholar]
- Poon E, Ellis DA, Fitzgerald HE, Zucker RA. Intellectual, cognitive, and academic performance among sons of alcoholics, during the early school years: differences related to subtypes of familial alcoholism. Alcohol Clin Exp Res. 2000;24:1020–1027. [PubMed] [Google Scholar]
- Pruss RM. Receptors for glutamate and other excitatory amino acids: A cause for excitement in nervous system development. In: Zagon I, McLaughlin P, editors. Receptors in the developing nervous system: Neurotransmitters. Vol. 2. New York: Chapman & Hall; 1993. pp. 141–162. [Google Scholar]
- Puts NA, Edden RA. In vivo magnetic resonance spectroscopy of GABA: a methodological review. Prog Nucl Magn Reson Spectrosc. 2012;60:29–41. doi: 10.1016/j.pnmrs.2011.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Puts NA, Edden RA, Evans CJ, McGlone F, McGonigle DJ. Regionally specific human GABA concentration correlates with tactile discrimination thresholds. Journal of Neuroscience. 2011;31:16556–16560. doi: 10.1523/JNEUROSCI.4489-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pyapali GK, Turner DA, Wilson WA, Swartzwelder HS. Age and dose-dependent effects of ethanol on the induction of hippocampal long-term potentiation. Alcohol. 1999;19:107–111. doi: 10.1016/s0741-8329(99)00021-x. [DOI] [PubMed] [Google Scholar]
- Rajendran P, Spear LP. The effects of ethanol on spatial and nonspatial memory in adolescent and adult rats studied using an appetitive paradigm. Ann N Y Acad Sci. 2004;1021:441–444. doi: 10.1196/annals.1308.060. [DOI] [PubMed] [Google Scholar]
- Ramirez RL, Spear LP. Ontogeny of ethanol-induced motor impairment following acute ethanol: assessment via the negative geotaxis reflex in adolescent and adult rats. Pharmacol Biochem Behav. 2010;95:242–248. doi: 10.1016/j.pbb.2010.01.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Raz N, Lindenberger U, Rodrigue KM, Kennedy KM, Head D, Williamson A, Dahle C, Gerstorf D, Acker JD. Regional brain changes in aging healthy adults: general trends, individual differences and modifiers. Cereb Cortex. 2005;15:1676–1689. doi: 10.1093/cercor/bhi044. [DOI] [PubMed] [Google Scholar]
- Raz N, Rodrigue KM, Haacke EM. Brain aging and its modifiers: insights from in vivo neuromorphometry and susceptibility weighted imaging. Ann N Y Acad Sci. 2007;1097:84–93. doi: 10.1196/annals.1379.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Riggs NR, Greenberg MT. Neurocognition as a moderator and mediator in adolescent substance misuse prevention. Am J Drug Alcohol Abuse. 2009;35:209–213. doi: 10.1080/00952990903005940. [DOI] [PubMed] [Google Scholar]
- Roberto M, Madamba SG, Stouffer DG, Parsons LH, Siggins GR. Increased GABA release in the central amygdala of ethanol-dependent rats. J Neurosci. 2004;24:10159–10166. doi: 10.1523/JNEUROSCI.3004-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rodd ZA, Bell RL, McKinzie DL, Webster AA, Murphy JM, Lumeng L, Li TK, McBride WJ. Low-dose stimulatory effects of ethanol during adolescence in rat lines selectively bred for high alcohol intake. Alcohol Clin Exp Res. 2004;28:535–543. doi: 10.1097/01.alc.0000122107.08417.d0. [DOI] [PubMed] [Google Scholar]
- Rodnick ME, Hockley BG, Sherman P, Quesada C, Battle MR, Jackson A, Linder KE, Macholl S, Trigg WJ, Kilbourn MR, Scott PJ. Novel fluorine-18 PET radiotracers based on flumazenil for GABAA imaging in the brain. Nucl Med Biol. 2013;40:901–905. doi: 10.1016/j.nucmedbio.2013.06.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosso IM, Weiner MR, Crowley DJ, Silveri MM, Rauch SL, Jensen JE. Insula and Anterior Cingulate Gaba Levels in Posttraumatic Stress Disorder: Preliminary Findings Using Magnetic Resonance Spectroscopy. Depress Anxiety. 2013 doi: 10.1002/da.22155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosso IM, Young AD, Femia LA, Yurgelun-Todd DA. Cognitive and emotional components of frontal lobe functioning in childhood and adolescence. Ann N Y Acad Sci. 2004;1021:355–362. doi: 10.1196/annals.1308.045. [DOI] [PubMed] [Google Scholar]
- Rothman DL, Petroff OAC, Behar KL, Mattson RH. Localized 1H NMR measurements of (gamma)-aminobutyric acid in human brain in vivo. Proceedings of the National Academy of Sciences of the United States of America. 1993;90:5662–5666. doi: 10.1073/pnas.90.12.5662. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rubia K, Overmeyer S, Taylor E, Brammer M, Williams SC, Simmons A, Andrew C, Bullmore ET. Functional frontalisation with age: mapping neurodevelopmental trajectories with fMRI. Neurosci Biobehav Rev. 2000;24:13–19. doi: 10.1016/s0149-7634(99)00055-x. [DOI] [PubMed] [Google Scholar]
- SAMHSA. NSDUH Series H-38A, HHS Publication No. SMA 10-4856Findings. Office of Applied Studies; Rockville, MD: 2010a. Results from the 2009 National Survey on Drug Use and Health: Volume I. Summary of National Findings. [Google Scholar]
- SAMHSA. The OAS Report: A Day in the Life of American Adolescents: Substance Use Facts Update. Rockville, MD: 2010b. Substance Abuse and Mental Health Services Administration, Office of Applied Studies. [PubMed] [Google Scholar]
- Sanacora G, Fenton LR, Fasula MK, Rothman DL, Levin Y, Krystal JH, Mason GF. Cortical gamma-aminobutyric acid concentrations in depressed patients receiving cognitive behavioral therapy. Biol Psychiatry. 2006;59:284–286. doi: 10.1016/j.biopsych.2005.07.015. [DOI] [PubMed] [Google Scholar]
- Sanacora G, Mason GF, Rothman DL, Hyder F, Ciarcia JJ, Ostroff RB, Berman RM, Krystal JH. Increased cortical GABA concentrations in depressed patients receiving ECT. Am J Psychiatry. 2003;160:577–579. doi: 10.1176/appi.ajp.160.3.577. [DOI] [PubMed] [Google Scholar]
- Sanacora G, Mason GF, Rothman DL, Krystal JH. Increased occipital cortex GABA concentrations in depressed patients after therapy with selective serotonin reuptake inhibitors. Am J Psychiatry. 2002;159:663–665. doi: 10.1176/appi.ajp.159.4.663. [DOI] [PubMed] [Google Scholar]
- Santerre JL, Gigante ED, Landin JD, Werner DF. Molecular and behavioral characterization of adolescent protein kinase C following high dose ethanol exposure. Psychopharmacology (Berl) 2013 doi: 10.1007/s00213-013-3267-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saransaari P, Oja SS. Dizocilpine binding to cerebral cortical membranes from developing and ageing mice. Mech Ageing Dev. 1995;85:171–181. doi: 10.1016/0047-6374(95)01665-1. [DOI] [PubMed] [Google Scholar]
- Schramm-Sapyta NL, Walker QD, Caster JM, Levin ED, Kuhn CM. Are adolescents more vulnerable to drug addiction than adults? Evidence from animal models. Psychopharmacology (Berl) 2009;206:1–21. doi: 10.1007/s00213-009-1585-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schuckit MA, Tapert S, Matthews SC, Paulus MP, Tolentino NJ, Smith TL, Trim RS, Hall S, Simmons A. fMRI differences between subjects with low and high responses to alcohol during a stop signal task. Alcohol Clin Exp Res. 2012;36:130–140. doi: 10.1111/j.1530-0277.2011.01590.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schulteis G, Archer C, Tapert SF, Frank LR. Intermittent binge alcohol exposure during the periadolescent period induces spatial working memory deficits in young adult rats. Alcohol. 2008;42:459–467. doi: 10.1016/j.alcohol.2008.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schumann G, Johann M, Frank J, Preuss U, Dahmen N, Laucht M, Rietschel M, Rujescu D, Lourdusamy A, Clarke TK, Krause K, Dyer A, Depner M, Wellek S, Treutlein J, Szegedi A, Giegling I, Cichon S, Blomeyer D, Heinz A, Heath S, Lathrop M, Wodarz N, Soyka M, Spanagel R, Mann K. Systematic analysis of glutamatergic neurotransmission genes in alcohol dependence and adolescent risky drinking behavior. Arch Gen Psychiatry. 2008;65:826–838. doi: 10.1001/archpsyc.65.7.826. [DOI] [PubMed] [Google Scholar]
- Schwandt ML, Barr CS, Suomi SJ, Higley JD. Age-dependent variation in behavior following acute ethanol administration in male and female adolescent rhesus macaques (Macaca mulatta) Alcohol Clin Exp Res. 2007;31:228–237. doi: 10.1111/j.1530-0277.2006.00300.x. [DOI] [PubMed] [Google Scholar]
- Schweinsburg AD, McQueeny T, Nagel BJ, Eyler LT, Tapert SF. A preliminary study of functional magnetic resonance imaging response during verbal encoding among adolescent binge drinkers. Alcohol. 2010;44:111–117. doi: 10.1016/j.alcohol.2009.09.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schweinsburg AD, Paulus MP, Barlett VC, Killeen LA, Caldwell LC, Pulido C, Brown SA, Tapert SF. An FMRI study of response inhibition in youths with a family history of alcoholism. Ann N Y Acad Sci. 2004;1021:391–394. doi: 10.1196/annals.1308.050. [DOI] [PubMed] [Google Scholar]
- Seilicovich A, Duvilanski B, Gonzalez NN, Rettori V, Mitridate de Novara A, Maines VM, Fiszer de Plazas S. The effect of acute ethanol administration on GABA receptor binding in cerebellum and hypothalamus. Eur J Pharmacol. 1985;111:365–368. doi: 10.1016/0014-2999(85)90643-0. [DOI] [PubMed] [Google Scholar]
- Shmuel A, Augath M, Oeltermann A, Logothetis NK. Negative functional MRI response correlates with decreases in neuronal activity in monkey visual area V1. Nat Neurosci. 2006;9:569–577. doi: 10.1038/nn1675. [DOI] [PubMed] [Google Scholar]
- Shmuel A, Yacoub E, Pfeuffer J, Van de Moortele PF, Adriany G, Hu X, Ugurbil K. Sustained negative BOLD, blood flow and oxygen consumption response and its coupling to the positive response in the human brain. Neuron. 2002;36:1195–1210. doi: 10.1016/s0896-6273(02)01061-9. [DOI] [PubMed] [Google Scholar]
- Sieghart W, Sperk G. Subunit composition, distribution and function of GABA(A) receptor subtypes. Curr Top Med Chem. 2002;2:795–816. doi: 10.2174/1568026023393507. [DOI] [PubMed] [Google Scholar]
- Sigel E, Steinmann ME. Structure, function, and modulation of GABA(A) receptors. J Biol Chem. 2012;287:40224–40231. doi: 10.1074/jbc.R112.386664. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silveri MM. Adolescent brain development and underage drinking in the United States: identifying risks of alcohol use in college populations. Harv Rev Psychiatry. 2012;20:189–200. doi: 10.3109/10673229.2012.714642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silveri MM, Cohen-Gilbert J, Crowley DJ, Rosso IM, Jensen JE, Sneider JT. Altered anterior cingulate neurochemistry in emerging adult binge drinkers with a history of alcohol-induced blackouts. Alcohol Clin Exp Res. 2014 doi: 10.1111/acer.12346. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silveri MM, Rogowska J, McCaffrey A, Yurgelun-Todd DA. Adolescents at risk for alcohol abuse demonstrate altered frontal lobe activation during stroop performance. Alcoholism, Clinical and Experimental Research. 2011;35:218–228. doi: 10.1111/j.1530-0277.2010.01337.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silveri MM, Sneider JT, Crowley DJ, Covell MJ, Acharya D, Rosso IM, Jensen JE. Frontal lobe gamma-aminobutyric acid levels during adolescence: associations with impulsivity and response inhibition. Biol Psychiatry. 2013;74:296–304. doi: 10.1016/j.biopsych.2013.01.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silveri MM, Spear LP. Decreased sensitivity to the hypnotic effects of ethanol early in ontogeny. Alcohol Clin Exp Res. 1998;22:670–676. doi: 10.1111/j.1530-0277.1998.tb04310.x. [DOI] [PubMed] [Google Scholar]
- Silveri MM, Spear LP. Ontogeny of rapid tolerance to the hypnotic effects of ethanol. Alcohol Clin Exp Res. 1999;23:1180–1184. [PubMed] [Google Scholar]
- Silveri MM, Spear LP. Acute, rapid, and chronic tolerance during ontogeny: observations when equating ethanol perturbation across age. Alcohol Clin Exp Res. 2001;25:1301–1308. [PubMed] [Google Scholar]
- Silveri MM, Spear LP. The effects of NMDA and GABAA pharmacological manipulations on ethanol sensitivity in immature and mature animals. Alcohol Clin Exp Res. 2002;26:449–456. [PubMed] [Google Scholar]
- Silveri MM, Spear LP. The effects of NMDA and GABAA pharmacological manipulations on acute and rapid tolerance to ethanol during ontogeny. Alcohol Clin Exp Res. 2004;28:884–894. doi: 10.1097/01.alc.0000128221.68382.ba. [DOI] [PubMed] [Google Scholar]
- Silveri MM, Tzilos GK, Pimentel PJ, Yurgelun-Todd DA. Trajectories of adolescent emotional and cognitive development: effects of sex and risk for drug use. Ann N Y Acad Sci. 2004;1021:363–370. doi: 10.1196/annals.1308.046. [DOI] [PubMed] [Google Scholar]
- Silveri MM, Tzilos GK, Yurgelun-Todd DA. Relationship between white matter volume and cognitive performance during adolescence: effects of age, sex and risk for drug use. Addiction. 2008;103:1509–1520. doi: 10.1111/j.1360-0443.2008.02272.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silvers JM, Tokunaga S, Mittleman G, Matthews DB. Chronic intermittent injections of high-dose ethanol during adolescence produce metabolic, hypnotic, and cognitive tolerance in rats. Alcohol Clin Exp Res. 2003;27:1606–1612. doi: 10.1097/01.ALC.0000090141.66526.22. [DOI] [PubMed] [Google Scholar]
- Silvers JM, Tokunaga S, Mittleman G, O’Buckley T, Morrow AL, Matthews DB. Chronic intermittent ethanol exposure during adolescence reduces the effect of ethanol challenge on hippocampal allopregnanolone levels and Morris water maze task performance. Alcohol. 2006;39:151–158. doi: 10.1016/j.alcohol.2006.09.001. [DOI] [PubMed] [Google Scholar]
- Simister RJ, McLean MA, Barker GJ, Duncan JS. Proton MRS reveals frontal lobe metabolite abnormalities in idiopathic generalized epilepsy. Neurology. 2003;61:897–902. doi: 10.1212/01.wnl.0000086903.69738.dc. [DOI] [PubMed] [Google Scholar]
- Simister RJ, McLean MA, Barker GJ, Duncan JS. Proton MR spectroscopy of metabolite concentrations in temporal lobe epilepsy and effect of temporal lobe resection. Epilepsy Res. 2009;83:168–176. doi: 10.1016/j.eplepsyres.2008.11.006. [DOI] [PubMed] [Google Scholar]
- Simpson HB, Shungu DC, Bender J, Jr, Mao X, Xu X, Slifstein M, Kegeles LS. Investigation of cortical glutamate-glutamine and gamma-aminobutyric acid in obsessive-compulsive disorder by proton magnetic resonance spectroscopy. Neuropsychopharmacology. 2012;37:2684–2692. doi: 10.1038/npp.2012.132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sircar R, Sircar D. Adolescent rats exposed to repeated ethanol treatment show lingering behavioral impairments. Alcohol Clin Exp Res. 2005;29:1402–1410. doi: 10.1097/01.alc.0000175012.77756.d9. [DOI] [PubMed] [Google Scholar]
- Slawecki CJ, Betancourt M, Cole M, Ehlers CL. Periadolescent alcohol exposure has lasting effects on adult neurophysiological function in rats. Brain Res Dev Brain Res. 2001;128:63–72. doi: 10.1016/s0165-3806(01)00150-x. [DOI] [PubMed] [Google Scholar]
- Slawecki CJ, Roth J. Comparison of the onset of hypoactivity and anxiety-like behavior during alcohol withdrawal in adolescent and adult rats. Alcohol Clin Exp Res. 2004;28:598–607. doi: 10.1097/01.alc.0000122767.69206.1b. [DOI] [PubMed] [Google Scholar]
- Slawecki CJ, Roth J, Gilder A. Neurobehavioral profiles during the acute phase of ethanol withdrawal in adolescent and adult Sprague-Dawley rats. Behav Brain Res. 2006;170:41–51. doi: 10.1016/j.bbr.2006.01.023. [DOI] [PubMed] [Google Scholar]
- Soghomonian JJ, Martin DL. Two isoforms of glutamate decarboxylase: why? Trends Pharmacol Sci. 1998;19:500–505. doi: 10.1016/s0165-6147(98)01270-x. [DOI] [PubMed] [Google Scholar]
- Sowell ER, Delis D, Stiles J, Jernigan TL. Improved memory functioning and frontal lobe maturation between childhood and adolescence: a structural MRI study. J Int Neuropsychol Soc. 2001;7:312–322. doi: 10.1017/s135561770173305x. [DOI] [PubMed] [Google Scholar]
- Sowell ER, Jernigan TL. Further MRI evidence of late brain maturation: Limbic volume increases and changing asymmetries during childhood and adolescence. Developmental Neuropsychology. 1998;14:599–617. [Google Scholar]
- Sowell ER, Thompson PM, Tessner KD, Toga AW. Mapping continued brain growth and gray matter density reduction in dorsal frontal cortex: Inverse relationships during postadolescent brain maturation. J Neurosci. 2001;21:8819–8829. doi: 10.1523/JNEUROSCI.21-22-08819.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sowell ER, Thompson PM, Toga AW. Mapping changes in the human cortex throughout the span of life. Neuroscientist. 2004;10:372–392. doi: 10.1177/1073858404263960. [DOI] [PubMed] [Google Scholar]
- Soyka M, Preuss UW, Hesselbrock V, Zill P, Koller G, Bondy B. GABA-A2 receptor subunit gene (GABRA2) polymorphisms and risk for alcohol dependence. J Psychiatr Res. 2008;42:184–191. doi: 10.1016/j.jpsychires.2006.11.006. [DOI] [PubMed] [Google Scholar]
- Spadoni AD, Norman AL, Schweinsburg AD, Tapert SF. Effects of family history of alcohol use disorders on spatial working memory BOLD response in adolescents. Alcohol Clin Exp Res. 2008;32:1135–1145. doi: 10.1111/j.1530-0277.2008.00694.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spear LP. The adolescent brain and age-related behavioral manifestations. Neurosci Biobehav Rev. 2000;24:417–463. doi: 10.1016/s0149-7634(00)00014-2. [DOI] [PubMed] [Google Scholar]
- Spear LP. Adolescent neurobehavioral characteristics, alcohol sensitivities, and intake: Setting the stage for alcohol use disorders? Child Dev Perspect. 2011a;5:231–238. doi: 10.1111/j.1750-8606.2011.00182.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spear LP. Rewards, aversions and affect in adolescence: emerging convergences across laboratory animal and human data. Dev Cogn Neurosci. 2011b;1:392–400. doi: 10.1016/j.dcn.2011.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spear LP, Varlinskaya EI. Adolescence. Alcohol sensitivity, tolerance, and intake. Recent Dev Alcohol. 2005;17:143–159. [PubMed] [Google Scholar]
- Spear LP, Varlinskaya EI. Sensitivity to ethanol and other hedonic stimuli in an animal model of adolescence: implications for prevention science? Dev Psychobiol. 2010;52:236–243. doi: 10.1002/dev.20457. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Squeglia LM, Pulido C, Wetherill RR, Jacobus J, Brown GG, Tapert SF. Brain response to working memory over three years of adolescence: influence of initiating heavy drinking. J Stud Alcohol Drugs. 2012;73:749–760. doi: 10.15288/jsad.2012.73.749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stagg CJ. Magnetic Resonance Spectroscopy as a tool to study the role of GABA in motor-cortical plasticity. Neuroimage. 2013 doi: 10.1016/j.neuroimage.2013.01.009. [DOI] [PubMed] [Google Scholar]
- Stagg CJ, Bachtiar V, Johansen-Berg H. The role of GABA in human motor learning. Current Biology. 2011a;21:480–484. doi: 10.1016/j.cub.2011.01.069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stagg CJ, Bachtiar V, Johansen-Berg H. What are we measuring with GABA magnetic resonance spectroscopy? Commun Integr Biol. 2011b;4:573–575. doi: 10.4161/cib.4.5.16213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stagg CJ, Lang B, Best JG, McKnight K, Cavey A, Johansen-Berg H, Vincent A, Palace J. Autoantibodies to glutamic acid decarboxylase in patients with epilepsy are associated with low cortical GABA levels. Epilepsia. 2010;51:1898–1901. doi: 10.1111/j.1528-1167.2010.02644.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Staley JK, Gottschalk C, Petrakis IL, Gueorguieva R, O’Malley S, Baldwin R, Jatlow P, Verhoeff NP, Perry E, Weinzimmer D, Frohlich E, Ruff E, van Dyck CH, Seibyl JP, Innis RB, Krystal JH. Cortical gamma-aminobutyric acid type A-benzodiazepine receptors in recovery from alcohol dependence: relationship to features of alcohol dependence and cigarette smoking. Arch Gen Psychiatry. 2005;62:877–888. doi: 10.1001/archpsyc.62.8.877. [DOI] [PubMed] [Google Scholar]
- Stevens MC, Kiehl KA, Pearlson GD, Calhoun VD. Functional neural networks underlying response inhibition in adolescents and adults. Behav Brain Res. 2007;181:12–22. doi: 10.1016/j.bbr.2007.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Streeter CC, Gerbarg PL, Saper RB, Ciraulo DA, Brown RP. Effects of yoga on the autonomic nervous system, gamma-aminobutyric-acid, and allostasis in epilepsy, depression, and post-traumatic stress disorder. Med Hypotheses. 2012;78:571–579. doi: 10.1016/j.mehy.2012.01.021. [DOI] [PubMed] [Google Scholar]
- Streeter CC, Hennen J, Ke Y, Jensen JE, Sarid-Segal O, Nassar LE, Knapp C, Meyer AA, Kwak T, Renshaw PF, Ciraulo DA. Prefrontal GABA levels in cocaine-dependent subjects increase with pramipexole and venlafaxine treatment. Psychopharmacology (Berl) 2005;182:516–526. doi: 10.1007/s00213-005-0121-5. [DOI] [PubMed] [Google Scholar]
- Streeter CC, Jensen JE, Perlmutter RM, Cabral HJ, Tian H, Terhune DB, Ciraulo DA, Renshaw PF. Yoga Asana sessions increase brain GABA levels: a pilot study. J Altern Complement Med. 2007;13:419–426. doi: 10.1089/acm.2007.6338. [DOI] [PubMed] [Google Scholar]
- Streeter CC, Whitfield TH, Owen L, Rein T, Karri SK, Yakhkind A, Perlmutter R, Prescot A, Renshaw PF, Ciraulo DA, Jensen JE. Effects of yoga versus walking on mood, anxiety, and brain GABA levels: a randomized controlled MRS study. J Altern Complement Med. 2010;16:1145–1152. doi: 10.1089/acm.2010.0007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sullivan EV, Pfefferbaum A. Neurocircuitry in alcoholism: a substrate of disruption and repair. Psychopharmacology (Berl) 2005;180:583–594. doi: 10.1007/s00213-005-2267-6. [DOI] [PubMed] [Google Scholar]
- Sullivan EV, Pfefferbaum A, Rohlfing T, Baker FC, Padilla ML, Colrain IM. Developmental change in regional brain structure over 7 months in early adolescence: comparison of approaches for longitudinal atlas-based parcellation. Neuroimage. 2011;57:214–224. doi: 10.1016/j.neuroimage.2011.04.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sumner P, Edden RA, Bompas A, Evans CJ, Singh KD. More GABA, less distraction: a neurochemical predictor of motor decision speed. Nat Neurosci. 2010;13:825–827. doi: 10.1038/nn.2559. [DOI] [PubMed] [Google Scholar]
- Suzuki M, Hagino H, Nohara S, Zhou SY, Kawasaki Y, Takahashi T, Matsui M, Seto H, Ono T, Kurachi M. Male-specific volume expansion of the human hippocampus during adolescence. Cereb Cortex. 2005;15:187–193. doi: 10.1093/cercor/bhh121. [DOI] [PubMed] [Google Scholar]
- Swartzwelder HS, Richardson RC, Markwiese-Foerch B, Wilson WA, Little PJ. Developmental differences in the acquisition of tolerance to ethanol. Alcohol. 1998;15:311–314. doi: 10.1016/s0741-8329(97)00135-3. [DOI] [PubMed] [Google Scholar]
- Swartzwelder HS, Wilson WA, Tayyeb MI. Age-dependent inhibition of long-term potentiation by ethanol in immature versus mature hippocampus. Alcohol Clin Exp Res. 1995;19:1480–1485. doi: 10.1111/j.1530-0277.1995.tb01011.x. [DOI] [PubMed] [Google Scholar]
- Tapert SF, Brown GG, Baratta MV, Brown SA. fMRI BOLD response to alcohol stimuli in alcohol dependent young women. Addict Behav. 2004a;29:33–50. doi: 10.1016/j.addbeh.2003.07.003. [DOI] [PubMed] [Google Scholar]
- Tapert SF, Brown GG, Kindermann SS, Cheung EH, Frank LR, Brown SA. fMRI measurement of brain dysfunction in alcohol-dependent young women. Alcohol Clin Exp Res. 2001;25:236–245. [PubMed] [Google Scholar]
- Tapert SF, Pulido C, Paulus MP, Schuckit MA, Burke C. Level of response to alcohol and brain response during visual working memory. J Stud Alcohol. 2004b;65:692–700. doi: 10.15288/jsa.2004.65.692. [DOI] [PubMed] [Google Scholar]
- Tapert SF, Schweinsburg AD. The human adolescent brain and alcohol use disorders. Recent Dev Alcohol. 2005;17:177–197. doi: 10.1007/0-306-48626-1_9. [DOI] [PubMed] [Google Scholar]
- Tapert SF, Schweinsburg AD, Barlett VC, Brown SA, Frank LR, Brown GG, Meloy MJ. Blood oxygen level dependent response and spatial working memory in adolescents with alcohol use disorders. Alcohol Clin Exp Res. 2004c;28:1577–1586. doi: 10.1097/01.alc.0000141812.81234.a6. [DOI] [PubMed] [Google Scholar]
- Thakkar MM, Engemann SC, Sharma R, Mohan RR, Sahota P. Sleep-wakefulness in alcohol preferring and non-preferring rats following binge alcohol administration. Neuroscience. 2010;170:22–27. doi: 10.1016/j.neuroscience.2010.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tokunaga S, Silvers JM, Matthews DB. Chronic intermittent ethanol exposure during adolescence blocks ethanol-induced inhibition of spontaneously active hippocampal pyramidal neurons. Alcohol Clin Exp Res. 2006;30:1–6. doi: 10.1111/j.1530-0277.2006.00020.x. [DOI] [PubMed] [Google Scholar]
- Tolliver GA, Samson HH. The influence of early postweaning ethanol exposure on oral self-administration behavior in the rat. Pharmacol Biochem Behav. 1991;38:575–580. doi: 10.1016/0091-3057(91)90016-u. [DOI] [PubMed] [Google Scholar]
- Torres JM, Ortega E. Alcohol intoxication increases allopregnanolone levels in female adolescent humans. Neuropsychopharmacology. 2003;28:1207–1209. doi: 10.1038/sj.npp.1300170. [DOI] [PubMed] [Google Scholar]
- Torres JM, Ortega E. Alcohol intoxication increases allopregnanolone levels in male adolescent humans. Psychopharmacology (Berl) 2004;172:352–355. doi: 10.1007/s00213-003-1662-0. [DOI] [PubMed] [Google Scholar]
- Trim RS, Simmons AN, Tolentino NJ, Hall SA, Matthews SC, Robinson SK, Smith TL, Padula CB, Paulus MP, Tapert SF, Schuckit MA. Acute ethanol effects on brain activation in low- and high-level responders to alcohol. Alcohol Clin Exp Res. 2010;34:1162–1170. doi: 10.1111/j.1530-0277.2010.01193.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Trimpop RM, Kerr JH, Kirkcaldy BD. Comparing personality constructs of risk-taking behavior. Pers Individ Dif. 1999;26:237–254. [Google Scholar]
- Tseng KY, O’Donnell P. D2 dopamine receptors recruit a GABA component for their attenuation of excitatory synaptic transmission in the adult rat prefrontal cortex. Synapse. 2007;61:843–850. doi: 10.1002/syn.20432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tyacke RJ, Lingford-Hughes A, Reed LJ, Nutt DJ. GABAB receptors in addiction and its treatment. Adv Pharmacol. 2010;58:373–396. doi: 10.1016/S1054-3589(10)58014-1. [DOI] [PubMed] [Google Scholar]
- Uematsu A, Matsui M, Tanaka C, Takahashi T, Noguchi K, Suzuki M, Nishijo H. Developmental trajectories of amygdala and hippocampus from infancy to early adulthood in healthy individuals. PLoS One. 2012;7:e46970. doi: 10.1371/journal.pone.0046970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ueno S, Harris RA, Messing RO, Sanchez-Perez AM, Hodge CW, McMahon T, Wang D, Mehmert KK, Kelley SP, Haywood A, Olive MF, Buck KJ, Hood HM, Blednov Y, Findlay G, Mascia MP. Alcohol actions on GABA(A) receptors: from protein structure to mouse behavior. Alcohol Clin Exp Res. 2001;25:76S–81S. doi: 10.1097/00000374-200105051-00014. [DOI] [PubMed] [Google Scholar]
- Uhart M, Weerts EM, McCaul ME, Guo X, Yan X, Kranzler HR, Li N, Wand GS. GABRA2 markers moderate the subjective effects of alcohol. Addict Biol. 2013;18:357–369. doi: 10.1111/j.1369-1600.2012.00457.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Skike CE, Novier A, Diaz-Granados JL, Matthews DB. The effect of chronic intermittent ethanol exposure on spatial memory in adolescent rats: the dissociation of metabolic and cognitive tolerances. Brain Res. 2012;1453:34–39. doi: 10.1016/j.brainres.2012.03.006. [DOI] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP. Acute effects of ethanol on social behavior of adolescent and adult rats: role of familiarity of the test situation. Alcohol Clin Exp Res. 2002;26:1502–1511. doi: 10.1097/01.ALC.0000034033.95701.E3. [DOI] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP. Acute ethanol withdrawal (hangover) and social behavior in adolescent and adult male and female Sprague-Dawley rats. Alcohol Clin Exp Res. 2004a;28:40–50. doi: 10.1097/01.ALC.0000108655.51087.DF. [DOI] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP. Changes in sensitivity to ethanol-induced social facilitation and social inhibition from early to late adolescence. Ann N Y Acad Sci. 2004b;1021:459–461. doi: 10.1196/annals.1308.064. [DOI] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP. Differences in the social consequences of ethanol emerge during the course of adolescence in rats: social facilitation, social inhibition, and anxiolysis. Dev Psychobiol. 2006a;48:146–161. doi: 10.1002/dev.20124. [DOI] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP. Ontogeny of acute tolerance to ethanol-induced social inhibition in Sprague-Dawley rats. Alcohol Clin Exp Res. 2006b;30:1833–1844. doi: 10.1111/j.1530-0277.2006.00220.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP. Chronic tolerance to the social consequences of ethanol in adolescent and adult Sprague-Dawley rats. Neurotoxicol Teratol. 2007;29:23–30. doi: 10.1016/j.ntt.2006.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP, Spear NE. Social behavior and social motivation in adolescent rats: role of housing conditions and partner’s activity. Physiol Behav. 1999;67:475–482. doi: 10.1016/s0031-9384(98)00285-6. [DOI] [PubMed] [Google Scholar]
- Varlinskaya EI, Spear LP, Spear NE. Acute effects of ethanol on behavior of adolescent rats: role of social context. Alcohol Clin Exp Res. 2001;25:377–385. [PubMed] [Google Scholar]
- Verhoeff NP, Petroff OA, Hyder F, Zoghbi SS, Fujita M, Rajeevan N, Rothman DL, Seibyl JP, Mattson RH, Innis RB. Effects of vigabatrin on the GABAergic system as determined by [123I]iomazenil SPECT and GABA MRS. Epilepsia. 1999;40:1433–1438. doi: 10.1111/j.1528-1157.1999.tb02016.x. [DOI] [PubMed] [Google Scholar]
- Vetter-O’Hagen CS, Spear LP. Hormonal and physical markers of puberty and their relationship to adolescent-typical novelty-directed behavior. Dev Psychobiol. 2012;54:523–535. doi: 10.1002/dev.20610. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vincent SL, Pabreza L, Benes FM. Postnatal maturation of GABA-immunoreactive neurons of rat medial prefrontal cortex. J Comp Neurol. 1995;355:81–92. doi: 10.1002/cne.903550110. [DOI] [PubMed] [Google Scholar]
- Wahle P. Differential regulation of substance P and somatostatin in Martinotti cells of the developing cat visual cortex. J Comp Neurol. 1993;329:519–538. doi: 10.1002/cne.903290408. [DOI] [PubMed] [Google Scholar]
- Wallner M, Hanchar HJ, Olsen RW. Low dose acute alcohol effects on GABA A receptor subtypes. Pharmacol Ther. 2006;112:513–528. doi: 10.1016/j.pharmthera.2006.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weber OM, Verhagen A, Duc CO, Meier D, Leenders KL, Boesiger P. Effects of vigabatrin intake on brain GABA activity as monitored by spectrally edited magnetic resonance spectroscopy and positron emission tomography. Magn Reson Imaging. 1999;17:417–425. doi: 10.1016/s0730-725x(98)00184-2. [DOI] [PubMed] [Google Scholar]
- Weissenborn R, Duka T. Acute alcohol effects on cognitive function in social drinkers: their relationship to drinking habits. Psychopharmacology (Berl) 2003;165:306–312. doi: 10.1007/s00213-002-1281-1. [DOI] [PubMed] [Google Scholar]
- White AM, Bae JG, Truesdale MC, Ahmad S, Wilson WA, Swartzwelder HS. Chronic-intermittent ethanol exposure during adolescence prevents normal developmental changes in sensitivity to ethanol-induced motor impairments. Alcohol Clin Exp Res. 2002a;26:960–968. doi: 10.1097/01.ALC.0000021334.47130.F9. [DOI] [PubMed] [Google Scholar]
- White AM, Truesdale MC, Bae JG, Ahmad S, Wilson WA, Best PJ, Swartzwelder HS. Differential effects of ethanol on motor coordination in adolescent and adult rats. Pharmacol Biochem Behav. 2002b;73:673–677. doi: 10.1016/s0091-3057(02)00860-2. [DOI] [PubMed] [Google Scholar]
- Williams BR, Ponesse JS, Schachar RJ, Logan GD, Tannock R. Development of inhibitory control across the life span. Dev Psychol. 1999;35:205–213. doi: 10.1037//0012-1649.35.1.205. [DOI] [PubMed] [Google Scholar]
- Wills TA, Knapp DJ, Overstreet DH, Breese GR. Differential dietary ethanol intake and blood ethanol levels in adolescent and adult rats: effects on anxiety-like behavior and seizure thresholds. Alcohol Clin Exp Res. 2008;32:1350–1360. doi: 10.1111/j.1530-0277.2008.00709.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wills TA, Knapp DJ, Overstreet DH, Breese GR. Sensitization, duration, and pharmacological blockade of anxiety-like behavior following repeated ethanol withdrawal in adolescent and adult rats. Alcohol Clin Exp Res. 2009;33:455–463. doi: 10.1111/j.1530-0277.2008.00856.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Windle M, Spear LP, Fuligni AJ, Angold A, Brown JD, Pine D, Smith GT, Giedd J, Dahl RE. Transitions into underage and problem drinking: developmental processes and mechanisms between 10 and 15 years of age. Pediatrics. 2008;121(Suppl 4):S273–289. doi: 10.1542/peds.2007-2243C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Winkelman JW, Buxton OM, Jensen JE, Benson KL, O’Connor SP, Wang W, Renshaw PF. Reduced brain GABA in primary insomnia: preliminary data from 4T proton magnetic resonance spectroscopy (1H-MRS) Sleep. 2008;31:1499–1506. doi: 10.1093/sleep/31.11.1499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Witt ED. Research on alcohol and adolescent brain development: opportunities and future directions. Alcohol. 2010;44:119–124. doi: 10.1016/j.alcohol.2009.08.011. [DOI] [PubMed] [Google Scholar]
- Woollacott MH. Effects of ethanol on postural adjustments in humans. Exp Neurol. 1983;80:55–68. doi: 10.1016/0014-4886(83)90006-7. [DOI] [PubMed] [Google Scholar]
- Wu PH, Poelchen W, Proctor WR. Differential GABAB Receptor Modulation of Ethanol Effects on GABA(A) synaptic activity in hippocampal CA1 neurons. J Pharmacol Exp Ther. 2005;312:1082–1089. doi: 10.1124/jpet.104.075663. [DOI] [PubMed] [Google Scholar]
- Xia Y, Poosch MS, Whitty CJ, Kapatos G, Bannon MJ. GABA transporter mRNA: in vitro expression and quantitation in neonatal rat and postmortem human brain. Neurochem Int. 1993;22:263–270. doi: 10.1016/0197-0186(93)90054-9. [DOI] [PubMed] [Google Scholar]
- Yan H, Li Q, Fleming R, Madison RD, Wilson WA, Swartzwelder HS. Developmental sensitivity of hippocampal interneurons to ethanol: involvement of the hyperpolarization-activated current, Ih. J Neurophysiol. 2009;101:67–83. doi: 10.1152/jn.90557.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yoon JH, Maddock RJ, Rokem A, Silver MA, Minzenberg MJ, Ragland JD, Carter CS. GABA concentration is reduced in visual cortex in schizophrenia and correlates with orientation-specific surround suppression. J Neurosci. 2010;30:3777–3781. doi: 10.1523/JNEUROSCI.6158-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu ZY, Wang W, Fritschy JM, Witte OW, Redecker C. Changes in neocortical and hippocampal GABAA receptor subunit distribution during brain maturation and aging. Brain Res. 2006;1099:73–81. doi: 10.1016/j.brainres.2006.04.118. [DOI] [PubMed] [Google Scholar]
- Yurgelun-Todd D. Emotional and cognitive changes during adolescence. Curr Opin Neurobiol. 2007;17:251–257. doi: 10.1016/j.conb.2007.03.009. [DOI] [PubMed] [Google Scholar]
- Zhang H, Wang W, Gao W, Ge Y, Zhang J, Wu S, Xu L. Effect of propofol on the levels of neurotransmitters in normal human brain: a magnetic resonance spectroscopy study. Neurosci Lett. 2009;467:247–251. doi: 10.1016/j.neulet.2009.10.052. [DOI] [PubMed] [Google Scholar]
- Zhang Z, Morse AC, Koob GF, Schulteis G. Dose- and time-dependent expression of anxiety-like behavior in the elevated plus-maze during withdrawal from acute and repeated intermittent ethanol intoxication in rats. Alcohol Clin Exp Res. 2007;31:1811–1819. doi: 10.1111/j.1530-0277.2007.00483.x. [DOI] [PMC free article] [PubMed] [Google Scholar]


