Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jun 1.
Published in final edited form as: Alcohol. 2012 Mar 24;46(4):317–327. doi: 10.1016/j.alcohol.2011.12.002

Effects of alcohol on the membrane excitability and synaptic transmission of medium spiny neurons in the nucleus accumbens

Vincent N Marty 1,*, Igor Spigelman 1
PMCID: PMC3586202  NIHMSID: NIHMS366311  PMID: 22445807

Abstract

Chronic and excessive alcohol drinking lead to alcohol dependence and loss of control over alcohol consumption, with serious detrimental health consequences. Chronic alcohol exposure followed by protracted withdrawal causes profound alterations in the brain reward system that leads to marked changes in reinforcement mechanisms and motivational state. These long-lasting neuroadaptations are thought to contribute to the development of cravings and relapse. The nucleus accumbens (NAcc), a central component of the brain reward system, plays a critical role in alcohol-induced neuroadaptive changes underlying alcohol-seeking behaviors. Here we review the findings that chronic alcohol exposure produces long-lasting neuroadaptive changes in various ion channels that govern intrinsic membrane properties and neuronal excitability, as well as excitatory and inhibitory synaptic transmission in the NAcc that underlie alcohol-seeking behavior during protracted withdrawal.

Keywords: Alcoholism, Nucleus accumbens, Medium spiny neuron, Glutamate, GABA

Introduction

The mesolimbic dopamine system is part of the motivational system that regulates appropriate behavioral responses to natural reinforcers such as food, drink and sex. Natural rewards rarely produce long-lasting changes in the mesolimbic system (Hyman, Malenka, & Nestler, 2006). However, alcohol like all other drugs of abuse acutely activates the mesolimbic dopamine system and, upon chronic exposure, produces long-lasting functional alterations in the brain reward system that leads to marked changes in reinforcement mechanisms and motivational state (Koob & Nestler, 1997; Wise, 1998). This persistent functional reorganization of the reward system defines a new equilibrium, different from homeostasis, which allows an apparent stability around a new ‘set-point’, termed allostasis (Le Moal, 2009). Allostatic mechanisms are involved in maintaining a functioning brain reward system, but at a price. For instance, persistent neuroadaptations induced by chronic alcohol exposure produce a negative affect state which contributes to craving for alcohol thereby enhancing the risk of relapse (Breese et al., 2005; Koob, 2003; Sinha, 2001).

The nucleus accumbens (NAcc) is a key structure of the mesolimbic dopaminergic rewarding system involved in behavioral effects of alcohol associated with dependence. Two functionally and anatomically distinct sub-regions, the core and the shell compose the NAcc. These NAcc sub-regions are specifically involved in alcohol-seeking and drug-seeking behaviors (Chaudhri, Sahuque, Cone, & Janak, 2008; Fuchs, Evans, Parker, & See, 2004; Fuchs, Ramirez, & Bell, 2008). A recent study has demonstrated that pharmacological inactivation of the NAcc core reduces conditioned responding to discrete alcohol cues, supporting an important role for the NAcc core in cue-induced relapse to alcohol-seeking (Chaudhri, Sahuque, Schairer, & Janak, 2010). In contrast, pharmacological inactivation of the NAcc shell attenuates the renewal of conditioned responding triggered by the alcohol context suggesting a role for the NAcc shell in context-induced relapse to alcohol-seeking (Chaudhri et al., 2010). The NAcc is mainly (∼95%) composed of GABAergic medium spiny neurons (MSNs) which project massively to the ventral pallidum, substantia nigra and the ventral tegmental area (VTA) (Chang & Kitai, 1985; Kalivas, Churchill, Klitenick, 1993; Nauta, Smith, Faulla, Domesick, 1978; Usuda, Tanaka, & Chiba, 1998; Xia et al., 2011). MSNs also form functional GABAergic interconnections within the NAcc (Taverna, van Dongen, Groenewegen, & Pennartz, 2004). MSNs receive extensive glutamatergic inputs from limbic areas such as prefrontal cortex, hippocampus and basolateral amygdala (Brog, Salyapongse, Deutch, & Zahm, 1993; Sesack & Grace, 2010). The remaining neurons in the NAcc are subpopulations of local circuit cholinergic and GABAergic interneurons (Bennett & Bolam, 1994; de Quidt & Emson, 1986; Hussain, Johnson, & Totterdell, 1996; Vincent & Johansson, 1983; Vincent et al., 1983; Zhou, Vincent, & Dani, 2002). In addition to the well-described dopaminergic input from the VTA which is crucial in mediating reward-related information (Beckstead, Domesick, & Nauta, 1979; Dahlstrom & Fuxe, 1965; Fallon & Moore, 1978; Swanson, 1982), GABAergic neurons from the VTA also send projections to the NAcc (Van Bockstaele & Pickel, 1995), thus providing a reciprocal inhibitory loop to this mesolimbic circuit. Several studies have shown that MSNs play a central role in mediating the reinforcing properties of alcohol (Nie, Rewal, Gill, Ron, & Janak, 2011; Rewal et al., 2011; Rewal et al., 2009). However, we still have a limited understanding of the molecular and cellular mechanisms involved in the effect of alcohol on MSNs that lead to alcohol-induced behavior. The purpose of this article is to provide an overview of the findings showing that acute and chronic alcohol (ethanol) exposure can cause neuroadaptive changes in neuronal excitability, glutamatergic and GABAergic synaptic transmission in the NAcc that underlie important alcohol-induced behaviors.

Alcohol exposure alters electrical membrane properties in NAcc MSNs

The regulation of the electrical membrane properties of MSNs is critical in the integration and processing of drug-related information converging on the NAcc. In vivo, MSNs display two membrane potential states: a hyperpolarized down-state and a more depolarized up-state during which action potential discharge occurs (O'Donnell & Grace, 1995; Wilson & Kawaguchi, 1996). Dysregulation of this two-state potential has been proposed as a major neuroadaptation underlying addiction (Kim, Park, Lee, Park, & Kim, 2011; Mu et al., 2010). Systemic administration of ethanol reduces glutamate-activated and spontaneously active MSNs in the NAcc of freely moving rats (Criado, Lee, Berg, & Henriksen, 1995, 1997). Although the action of acute ethanol has largely been studied, the mechanisms involved in this action of ethanol on MSN firing are still not yet fully understood.

Action of ethanol on voltage-gated ion channels

Activity of MSNs is regulated by the electrical membrane properties that depend on properties of voltage-gated ion channels involved in the generation of action potential (AP), including voltage-dependent Na+, Ca2+ and K+ channels (Nisenbaum, Xu, & Wilson, 1994; Steephen & Manchanda, 2009). Several studies have shown that brain expression of mRNAs coding for multiple voltage-gated ionic channels varies between lines of rodents known to differ markedly in voluntary alcohol consumption (Mulligan et al., 2006; Mulligan et al., 2011; Ponomarev et al., 2006; Tabakoff et al., 2008). For instance, it has been shown that the levels of alcohol consumption was positively correlated with the expression levels of the mRNA coding for the Na+ channel β4 subunit (scn4b), an auxiliary subunit involved in the regulation of neuronal activity (Grieco, Malhotra, Chen, Isom, & Raman, 2005), in several brain regions, including the NAcc (Mulligan et al., 2006; Tabakoff et al., 2008). This suggests that high voluntary alcohol consumption could be due at least in part to the reduction of the inhibitory effect of ethanol on Na+ channels produced by the high levels of expression of scn4b. Indeed, acute application of high concentration of ethanol (100–200 mM) inhibits voltage-gated Na+ channel functions expressed into Xenopus oocytes (Horishita & Harris, 2008; Shiraishi & Harris, 2004) and in dorsal root ganglion neurons (Wu & Kendig, 1998). In addition, the presynaptic effects of ethanol on glutamatergic and GABAergic synaptic transmission of dopamine neurons in the VTA have been shown to depend on tetrodotoxin-sensitive Na+ channels (Deng, Li, Zhou, & Ye, 2009; Xiao et al., 2009; Xiao & Ye, 2008). All together, these studies suggest that voltage-gated Na+ channels play an important role in the processing of alcohol-related information. Thus, further studies will be needed in order to investigate the mechanisms involved in the effects of ethanol on Na+ channels of MSNs in the NAcc.

The inhibition of L-type voltage-gated Ca2+ channel (VGCCs) by acute ethanol has been described in several preparations (Calton, Wilson, & Moore, 1999; Carlen et al., 1993; Hendricson, Thomas, Lippmann, & Morrisett, 2003; Walter & Messing, 1999; Zucca & Valenzuela, 2010). Alterations in L-type VGCC number and activity have been related tochronic ethanol effects and withdrawal hyperexcitability (Dolin, Little, Hudspith, Pagonis, & Littleton, 1987; Guppy, Crabbe, & Littleton, 1995; Guppy & Littleton, 1994; Perez-Velazquez, Valiante, & Carlen, 1994). Behavioral experiments have demonstrated that L-type VGCC inhibitors attenuate the signs of the alcohol withdrawal syndrome and reduce the neuronal hyper-excitability observed after withdrawal from chronic alcohol (Bailey, Molleman, & Little, 1998; Smith, Watson, Stephens, & Little, 1999; Veatch & Gonzalez, 2000; Watson & Little, 2002; Whittington & Little, 1991). In the NAcc, it has been shown that L-type VGCC activity is increased during acute ethanol withdrawal (Rossetti, Isola, De Vry, & Fadda, 1999). Interestingly, inhibition of L-type VGCCs attenuates certain behavioral/neurological signs associated with acute ethanol withdrawal by reversing the depletion of extracellular dopamine levels in the NAcc of rats withdrawn from ethanol (Rossetti et al., 1999). Further studies will be needed to characterize the long-lasting effects of ethanol on VGCCs in the NAcc to determine their involvement in alcohol relapse.

The current–voltage relationship of MSNs is characterized by an inward rectification detectable at hyperpolarized membrane potentials (Belleau & Warren, 2000; O'Donnell & Grace, 1993). This inward rectification has been attributed to the activation of voltage-dependent inwardly rectifying K+ (KIR) channels (Belleau & Warren, 2000). KIR channels are the major determinants of the input resistance and the hyperpolarized resting membrane potential of MSNs during the down-state (Nisenbaum & Wilson, 1995). Interestingly, acute ethanol drinking modulates the expression of mRNAs coding for the KIR channel subunits in the NAcc of alcohol-preferring mice (Mulligan et al., 2011). In vitro, acute application of ethanol induces a hyperpolarization associated with a decrease in input resistance of MSNs in the striatum (Blomeley, Cains, Smith, & Bracci, 2011). This ethanol-induced hyperpolarization is abolished in the presence of barium suggesting that the effect of ethanol may be due to an increase in potassium conductance through barium-sensitive KIR channels (Blomeley et al., 2011). We showed that chronic intermittent ethanol (CIE) exposure followed by 40 days of withdrawal produced an increase in the inward rectification of the current-voltage relationship associated with a decrease in the input resistance of MSNs in the NAcc core (Marty & Spigelman, 2011). Moreover, the resting membrane potential of MSNs was more hyperpolarized in the CIE-treated animals (Marty & Spigelman, 2011). These results suggest that CIE treatment could cause a long-lasting increase in KIR channel activity which could be critical in the state transition and synaptic integration of MSNs in NAcc core (Steephen & Manchanda, 2009).

Action of ethanol on BK and SK channels

Large- (BK) and small-conductance (SK) Ca2+-activated K+ channels represent one of the several important pharmacological targets of ethanol (Brodie, Scholz, Weiger, & Dopico, 2007; Mulholland, Becker, Woodward, & Chandler, 2011; Mulholland et al., 2009; Treistman & Martin, 2009). Acute ethanol has been found to increase BK and decrease SK channel activity in several preparations (Brodie, McElvain, Bunney, & Appel,1999; Davies et al., 2003; Dopico, Lemos, & Treitman, 1996; Dreixler, Jenkins, Cao, Roizen, & Houamed, 2000).

BK channels are potent regulators of spike repolarization, firing frequency and fast afterhyperpolarization (fAHP) in MSNs of the NAcc (Ishikawa et al., 2009). By speeding up spike repolarization, BK channels can facilitate high-frequency action potential discharge in response to excitatory synaptic input (Gu, Vervaeke, & Storm, 2007). In genetically modified mice, BK subunit composition influences the response to ethanol and controls the development of ethanol tolerance at the molecular, cellular and behavioral levels (Martin et al., 2008; Treistman & Martin, 2009). In the striatum, acute ethanol exposure induces a rapid and selective degradation of BK mRNA, via an epigenetic mechanism involving microRNA miR-9, resulting in a reorganization of BK mRNA splice variants encoding for BK channel isoforms with different intrinsic properties and sensitivity to ethanol (Pietrzykowski et al., 2008). In the NAcc, acute application of ethanol potentiates the open probability of somatic BK channels (Martin et al., 2004). In a recent study, it has been shown that chronic ethanol exposure induces a persistent tolerance of BK channels to acute ethanol in MSNs which is dependent upon the duration of previous ethanol exposure (Velazquez-Marrero et al., 2011). This longer-persistence tolerance is thought to be due to a reorganization of BK channel subunits resulting in increased expression of an alcohol-resistant BK channel isoform called STREX (stress axis-regulated exon) (Velazquez-Marrero al., 2011). STREX channels are more sensitive to calcium indicating that these channels have higher levels of activity and larger BK current for a given calcium concentration. Interestingly, our findings have shown that CIE treatment induced a faster repolarization of the AP and a potentiation of the fAHP amplitude in NAcc MSNs (Marty & Spigelman, 2011). All together, these studies suggest that alcohol-induced reorganization of BK channel isoforms could be critical in the development of tolerance, dependency and addiction to alcohol.

It has been shown recently that chronic exposure of self-administrated ethanol followed by 3 weeks of withdrawal decreased SK channel function resulting in decreased slow AHP amplitude and increased excitability of MSNs in the NAcc core (Hopf et al., 2010). In addition, in vivo infusion of SK channel activator into the NAcc core selectively reduced motivation to obtain alcohol after withdrawal. Interestingly, chlorzoxazone, a US Food and Drug Administration (FDA)-approved positive modulator of Ca2+-activated K+ channels (Cao, Dreixler, Roizen, Roberts, & Houamed, 2001; Liu, Lo, & Wu, 2003), reduces alcohol-seeking behavior in a rat model of CIE exposure (Hopf et al., 2011). By contrast, CIE treatment administrated by oral intubation did not produce any modification of the slow AHP and firing rate of MSNs in the NAcc core (Marty & Spigelman, 2011).

The relative dose, duration and the mode of administration of ethanol represent important criteria in the interpretation and comparison of the results between studies, as this could explain the apparent discrepancies between findings. Furthermore, the composition of the intrapipette solutions used for electrophysiological recordings could also explain the differences in the experimental results of different studies. For instance, the addition of high concentration of EGTA (ethylene glycol tetraacetic acid), a calcium chelator, in the intrapipette solution could mask a possible effect of ethanol on ion channels that are activated by intracellular calcium. Therefore, all these different parameters should be carefully considered when comparing the effects of ethanol from different studies in order to avoid misleading interpretations.

All together, these results suggest that chronic alcohol exposure induces long-lasting alterations of voltage-gated ion channels which regulate the electrical membrane properties of MSNs. These persistent neuroadaptations could lead to the alteration of processing and integration of information converging on the NAcc which may have important implications for alcohol-related behaviors.

Alcohol exposure alters synaptic glutamatergic transmission in the NAcc

The MSNs receive major excitatory glutamatergic inputs from different part of the brain such as the hippocampus, prefrontal cortex, amygdala and thalamus (Brog et al., 1993; Sesack & Grace, 2010). The regulation of glutamatergic transmission is critical for the maintenance of homeostasis of the reward system. The neuroadaptive changes of synaptic glutamatergic transmission have been shown to play an important role in the development of addictive behaviors (Kalivas, Lalumiere, Knackstedt, & Shen, 2009). Indeed, pharmacological inhibition of AMPA receptor (AMPAR) and NMDA receptor (NMDAR) has been shown to attenuate reinstatement of alcohol-seeking behavior (Backstrom & Hyytia, 2004; Sanchis-Segura et al., 2006). Therefore, alteration of glutamatergic transmission into the NAcc could induce profound changes in the reward system leading to the expression of drug-related disorders characteristic of the allostatic state (David, Ansseau, & Abriani, 2005; Kalivas, 2009; Stuber et al., 2011).

Acute effects of ethanol on synaptic glutamatergic transmission in the NAcc

Acutely, ethanol has been shown to alter glutamate transmission within the NAcc. Alcohol reduces presynaptic release of glutamate within the NAcc and inhibits glutamate postsynaptic receptor function (Nie, Madmba, & Siggins, 1994; Zhang, Hendricson, & Morrisett, 2005). NMDAR-mediated currents are more sensitive to the inhibitory effect of low concentrations of ethanol, whereas AMPAR-mediated currents are only affected by high concentration of ethanol (Nie et al., 1994). Acute ethanol vapor exposure can alter the expression of glutamatergic synaptic plasticity by blocking the induction of NMDAR-dependent long-term depression (LTD) of excitatory postsynaptic currents (EPSCs) elicited in MSNs from the NAcc shell (Jeanes, Buske, & Morrisett, 2011). This acute effect of ethanol is thought to be due to the potent inhibitory action of ethanol on GluN2B-containing NMDAR function in the NAcc (Maldve et al., 2002; Zhang et al., 2005). In addition, this inhibition of NMDAR function by acute ethanol has been shown to be responsible for the ethanol-induced attenuation of the NMDAR-dependent long-term potentiation (LTP) of EPSCs in several brain structures including the striatum (Izumi, Nagashima, Murayama, & Zorumski, 2005; Morrisett & Swartzwelder, 1993; Schummers & Browning, 2001; Tokuda, Izumi, & Zorumski, 2011; Weitlauf, Egli, Grueter, & Winder, 2004; Xie et al., 2009; Yin, Park, Adermark, & Lovinger, 2007). By contrast, chronic alcohol exposure produces an upregulation of NMDAR function and subunit expression in the mesolimbic system during withdrawal from alcohol (Lack, Diaz, Chappell, DuBois, McCool, 2007; Szumlinski, Ary, Lominac, Klugmann, & Kippin, 2008; Wang et al., 2007; Wang, Lanfranco, Gibb, & Ron, 2011). Indeed, CIE treatment has recently been shown to facilitate the LTP induction of NMDAR-mediated EPSCs in VTA dopaminergic neurons during short-term withdrawal (Bernier, Whitaker, & Morikawa, 2011). In addition, chronic ethanol exposure has been reported to facilitate the induction of NMDAR-dependent LTP in hippocampal CA1 neurons, demonstrating that acute and chronic ethanol exposure differentially affect the induction of LTP (Fujii, Yamazaki, Sugihara, & Wakabayashi, 2008). These results suggest that increased NMDAR function may be, at least in part, responsible for the neuronal hyperexcitability observed during withdrawal from chronic ethanol treatment (Follesa & Ticku, 1995; Hendricson et al., 2007; Kumari & Ticku, 2000).

Ethanol affects the regulation of extracellular glutamate homeostasis in the NAcc

NAcc extracellular glutamate is an important determinant of ethanol drinking. Acute and chronic ethanol administration has been shown to increase extracellular levels of glutamate in the NAcc (Dahchour, Hoffmanm, Deitrich, & de Witte, 2000; Lallemand, Ward, De Witte, & Verbanck, 2011; Melendez, Hicks, Cagle, & Kalivas, 2005; Szumlinski et al., 2007) (Table 1). Experimental manipulations that elevate or lower NAcc extracellular glutamate levels in the NAcc can modulate ethanol drinking behavior by increasing or decreasing ethanol consumption, respectively (Kapasova & Szumlinski, 2008; Szumlinski et al., 2008). In addition, chronic alcohol administration sensitizes the response to an acute alcohol challenge to elevate extracellular glutamate levels in the NAcc (Szumlinski et al., 2007). Consistently, a recent study has shown that an increase in NAcc core extracellular glutamate levels is associated with cue-induced reinstatement of alcohol-seeking behavior (Gass, Sinclair, Cleva, Widholm, & Olive, 2011). This alcohol-induced dysregulation of extracellular glutamate levels could be due to an alteration in the mechanism of release and re-uptake of glutamate as shown in cocaine studies (Baker et al., 2003; Pierce, Bell, Duffy, & Kalivas, 1996). In the NAcc core, basal extracellular glutamate is mainly derived from constitutive cystine/glutamate (cys/glu) exchange (Baker, Xi, Shen, Swanson, & Kalivas, 2002). The release of cytoplasmic glutamate is linked to cystine uptake and occurs more in glial cells than neurons. Extracellular glutamate levels are also regulated by glutamate transporters (Danbolt, 2001), in particular the glial transporter GLT1 located at the vicinity of the synaptic cleft. GLT1 allows to buffer glutamate released from presynaptic vesicle and via cys/glu exchanger (Pendyam, Mohan, Kalivas, & Nayar, 2009). In addition, the deletion of the Homer2 gene, a member of the Homer family involved in the regulation of glutamate signaling in the postsynaptic density, has been shown to reduce the levels of glutamate in the NAcc (Szumlinski et al., 2005). This regulation of extracellular glutamate levels allows for tonic stimulation of metabotropic glutamate receptors which, in turn, control synaptic glutamatergic transmission. Indeed, it has been shown that activation of cys/glu exchanger in the NAcc stimulates presynaptic metabotropic glutamate receptor 2/3 (mGlu2/3) activation, thereby decreasing the synaptic glutamate release probability (Moran, McFarland, Melendez, Kalivas, & Seamans, 2005). Self-administration of cocaine reduces the expression of both cys/glu exchanger and GLT1 in the NAcc core (Knackstedt, Melendez, & Kalivas, 2010; Madayag et al., 2007). In addition, repeated cocaine administration reduces the expression of Homer1 in the NAcc during protracted withdrawal (Swanson, Baker, Carson, Worley, & Kalivas, 2001). Consequently, repeated cocaine exposure strongly reduces the basal levels of extracellular glutamate leading to an increase in synaptic release of glutamate via decreased tonic activation of release-regulating mGlu2/3 and mGlu1/5 in the NAcc core (Madayag et al., 2007; Miguens et al., 2008; Pierce et al., 1996; Swanson et al., 2001; Xi, Baker, Shen, Carson, & Kalivas, 2002). Activation of cys/glu exchanger and GLT1 has been shown to restore basal extracellular glutamate levels in the NAcc core and prevent cue-induced reinstatement of cocaine seeking (Baker et al., 2003; Knackstedt et al., 2010; Moussawi et al., 2009; Sondheimer & Knackstedt, 2011). Consistently, stimulation of mGlu2/3 to reduce glutamate release has been shown to attenuate cocaine seeking (Baptista, Martin-Fardon, & Weiss, 2004; Peters & Kalivas, 2006). Interestingly, reduction in glutamatergic transmission via activation of mGlu2/3 or inhibition of postsynaptic mGlu5 attenuates alcohol-seeking behavior (Adams, Short, & Lawrence, 2010; Backstrom & Hyytia, 2005; Johnson et al., 2007). These findings suggest that like cocaine, alcohol produces similar neuroadaptive changes of synaptic glutamatergic transmission in the NAcc which lead to the development of alcohol-seeking behavior.

Table 1.

Effects of ethanol on the extracellular levels of glutamate in the NAcc.

References Animal model Ethanol treatment Measurement period Levels of extracellular glutamate in the NAcc
Dahchour et al., 2000 Male HAS (High-alcohol sensitivity) and LAS (Low-alcohol sensitivity) rats Acute i.p. injection of EtOH 2 g/kg Up to 3 h after EtOH injection ↗ from 100 to 180 min after injection
Melendez et al., 2005 Male Sprague- Dawley rats Chronic EtOH exposure: EtOH l g/kg i.p. daily for 7 days 24 h and 14 days after the last EtOH injection ↗ at 24 h. = at 14 days.
Note: No change in GLT1 protein expression at 24h
Sziimlinski et al., 2007 Male alcohol- preferring C57BL/6J (B6) mice Scheduled High Alcohol Consumption (SHAC): 6 doses of 5% EtOH (v/v) solution every other day (SHAC6) Day 1 for SHAC1 (1st dose of EtOH), and day 18 for SHAC6 (6th dose of EtOH) Basal levels SHAC1 = SHAC6.
↗ during EtOH consumption only in SHAC6.
↗ after acute EtOH challenge (2 g/kg, i.p) only in SHAC6
Kapasova & Szumlinski, 2008 Male alcohol- preferring C57BL/6J (B6) mice Male alcohol- avoiding DBA2/J (D2) mice Chronic intermittent EtOH exposure: 1–8 doses of EtOH 2 g/kg i.p. every other day 1st and 8th EtOH injection Basal levels = prior the 1st and 8th EtOH injection in both B6 and D2 mice.
↗ after the 1st EtOH injection only in D2 mice.
↗ after the 8th EtOH injection only in B6 mice.
Notes: APDC (mGlu2/3 agonist) ↘ EtOH intake in B6 and D2; TBOA (glutamate reuptake inhibitor) ↗ EtOH intake only in B6 mice
Szumlinski et al., 2008 Male alcohol- preferring C57BL/6J (B6) mice Chronic intermittent EtOH exposure: 8 doses of EtOH 2 g/kg i.p. every other day After acute i.p. injection of EtOH 2 g/kg ↗ in control mice.
↗↗ in mice overexpressing Homer2 in the NAcc.
Note: overexpression of Homer2 in the NAcc ↗ EtOH intake
Lallemand et al., 2011 Male Wistar rats Chronic intermittent ethanol exposure: EtOH 2 or 3 g/kg administrated by gavage 24 h after the last dose of EtOH Basal levels ↘ only in EtOH 2 g/kg group.
↗ after acute EtOH challenge in both groups
Gass et al., 2011 Male Wistar rats EtOH self- administration exposure: EtOH l mg/kg i.v. on a FR1 schedule During reinstatment testing session ↗ during cue-induced reinstatement of EtOH-seeking behavior.

Lasting effects of chronic ethanol exposure on synaptic glutamatergic transmission in the NAcc

Extensive studies using different ethanol exposure models have examined the effect of chronic ethanol exposure during withdrawal on glutamatergic receptor function in several brain regions including the mesolimbic dopamine system (Stuber, Hopf, Tye, Chen, & Bonci, 2010). In the NAcc of rats, it has been shown recently that CIE exposure induces the activation of AKT kinase and its intracellular target named mammalian target of rapamycin complex 1 (mTORC1) kinase (Neasta, Ben Hamida, Yowell, Carnicella, & Ron, 2010). mTORC1-related signaling plays an important role in the translocation of synaptic proteins (Hoeffer & Klann, 2010). Activation of the mTORC1-related pathway is thought to be responsible for the increased expression level of AMPAR subunit GluA1 and Homer in the NAcc observed at 24 h withdrawal after 3 months of CIE exposure (Neasta et al., 2010). Interestingly, inhibition of AKT or mTORC1 reduces alcohol-seeking behavior (Neasta et al., 2010; Neasta, Ben Hamida, Yowell, Carnicella, & Ron, 2011). Further, it has been shown that the deletion of the Homer2 gene impairs the development of behavioral and neurochemical plasticity after repeated ethanol administration (Szumlinski et al., 2005). In agreement, over-expression of Homer2 protein in the NAcc facilitates ethanol reinforcement and increases ethanol intake (Szumlinski et al., 2008). These studies suggest that mTORC1 and Homer2 activation may be involved in the altered expression of proteins involved in glutamatergic transmission after CIE treatment. For instance, Obara et al. (2009) have found that NMDAR subunits GluN2A and GluN2B, mGlu5 and mGlu1 protein expression increase in the NAcc core of rats after 24 h withdrawal following 6 months of CIE exposure. However, after 4 weeks of withdrawal only mGlu5 expression remains increased (Obara et al., 2009). Interestingly, the therapeutic efficacy of acamprosate, a US FDA-approved drug used for treating human alcoholism, is thought to relate to the inhibition of mGlu5 function (Blednov & Harris, 2008). These findings suggest that mGlu5 could be involved in mechanisms mediating alcohol-seeking behavior.

There is additional evidence that chronic ethanol exposure and withdrawal produces a time-dependent modulation of glutamatergic receptor function in the NAcc. Indeed, it has been recently demonstrated that the NMDAR-dependent LTD of EPSCs induced in MSNs from the NAcc shell is abolished after 24 h withdrawal of chronic ethanol vapor exposure, instead resulting in LTP of EPSCs. However, after 72 h of withdrawal when the NMDAR- dependent LTD was still strongly reduced, the LTP was no longer observed (Jeanes et al., 2011). All together, these studies suggest that the neuroadaptive changes observed during early withdrawal may lead to distinct long-term neuroadaptations involving different neuronal mechanisms that are more likely to underlie alcohol-seeking behaviors.

Increasing evidence suggests similarities between mechanisms induced by drugs of abuse (Hyman et al., 2006; Kalivas et al., 2009; Nestler, 2001). Cocaine, alcohol and other drugs of abuse cause long-lasting neuroadaptations of glutamatergic system in the brain including the NAcc (Gass & Olive, 2008). Protracted withdrawal from cocaine induces a long-lasting potentiation of AMPAR EPSC amplitude in the NAcc core that plays an important role in cue-induced cocaine-seeking behavior (Conrad et al., 2008; Mameli et al., 2009). This potentiation of AMPAR synaptic transmission is due to an increase in the number of postsynaptic Ca2+-permeable GluA1 homomeric (GluA2-lacking) AMPARs which is characterized by an increase in single channel conductance (Conrad et al., 2008; Guire, Oh, Soderling, & Derkach, 2008; Mameli et al., 2009). Our recent findings show that CIE exposure followed by prolonged withdrawal (>40 days) induced a potentiation of AMPAR synaptic transmission associated with an increase in AMPAR unitary conductance in the NAcc core (Marty & Spigelman, 2011). These results suggest that the potentiation of glutamate transmission induced during protracted withdrawal from CIE treatment could be due to a redistribution of AMPAR subunit composition increasing the number of functional postsynaptic GluA2-lacking AMPARs as previously observed in cocaine studies (Conrad et al., 2008; Mameli, Balland, Lujan, & Luscher, 2007; Mameli et al., 2009). Thus, increased NAcc core glutamatergic transmission may be a critical neuroadaptation that could be responsible for cue-induced reinstatement of alcohol-seeking behavior, which in abstaining human alcoholics would promote alcohol craving and increase the risk of relapse.

Alcohol exposure alters synaptic GABAergic transmission in the NAcc

The MSNs receive strong inhibitory GABAergic inputs from the VTA (Sesack & Grace, 2010; Van Bockstaele & Pickel, 1995; Voorn, Jorritsma-Byham, Van Dijk, & Bujis, 1986). Moreover, MSNs from the NAcc send reciprocal GABAergic projections back to non-dopaminergic neurons in the VTA (Kalivas et al., 1993; Usuda et al., 1998; Xia et al., 2011). This reciprocal connection between the NAcc and the VTA is thought to mediate an inhibitory feedback to regulate dopamine release in the NAcc (Rahman & McBride, 2000, 2002). In addition, the GABAergic interneurons within the NAcc regulate the excitability of the MSNs (Taverna, Canciani, & Pennartz, 2007). Therefore, modulation of GABAergic transmission in the NAcc is important in the regulation of the reward system homeostasis.

Acute effects of ethanol on synaptic GABAergic transmission in the NAcc

GABAA receptors (GABAARs) represent one of the several important pharmacological targets of ethanol (Kumar et al., 2009; Lovinger, 1997; Weiner & Valenzuela, 2006). Modulation of GABAergic transmission in the central nervous system underlies many of the behavioral effects of ethanol, including sedative-hypnotic, anticonvulsant, cognitive-impairing and anxiolytic actions (Kumar et al., 2009). Alcohol tolerance and dependence also appear to be attributable, at least in part, to changes in the function of GABAAR subunit composition and trafficking (Kumar, Fleming, & Morrow, 2004; Liang et al., 2007; Liang et al., 2006). Numerous studies have demonstrated that acute ethanol potentiates GABAAR function (Weiner & Valenzuela, 2006). In the NAcc core, ethanol (44–100 mM) increases GABA currents in about 50% of MSNs recorded (Nie, Madamba, & Siggins, 2000). This ethanol-induced enhancement of GABA currents is blocked by the non-selective mGlu receptor antagonist MCPG (Nie et al., 2000). In addition, it has been shown that the inhibitory effect of ethanol on NMDAR-mediated synaptic transmission in the NAcc core depends in part on the activation of metabotropic GABAB receptors (GABABRs) likely via a presynaptic mechanism (Steffensen, Nie, Criado, & Siggins, 2000). Interestingly, acamprosate, used therapeutically to reduce relapse, has the opposite effect of ethanol in the NAcc core. Acamprosate enhances NMDAR-mediated synaptic transmission and blocks presynaptic GABABRs (Berton, Franscesconi, Madamba, Zieglgansberger, & Siggins, 1998). Yet, baclofen, a muscle relaxant and agonist at GABABRs, is also very effective to suppress the alcohol withdrawal syndrome severity in human alcoholics with minimal side effects (Addolorato et al., 2006; Liu & Wang, 2011). Together these studies suggest that ethanol can modulate the intricate and reciprocal regulation of the GABAergic and glutamatergic transmission in the NAcc core that can lead to long-lasting neuroadaptations responsible for alcohol dependence and alcohol-seeking behavior.

Role of GABAergic receptor subunits of the NAcc in alcohol drinking behavior

Alcohol drinking behavior has been related to GABAergic transmissionin the NAcc. Morphometric studies have shown that alcohol preference is associated with increased density of GABAergic terminals in the NAcc (Hwang, Lumeng, Wu, & Li, 1990). Down-regulation of the α4-subunit of GABAAR has been observed in the NAcc during early withdrawal from chronic ethanol exposure (Papadeas, Grobin, & Morrow, 2001). Such early decreases in GABAAR α4-subunit expression are thought to underlie the early to lerance for alcohol (Liang et al., 2007). Recent studies have provided the first compelling evidence for a specific role of GABAARα4- and δ-subunits in mediating alcohol-drinking behavior (Nie et al., 2011; Rewal et al., 2011; Rewal et al., 2009). In an elegant series of studies, Janak and colleagues have demonstrated that selective reduction of the GABAAR α4-subunit expressions specifically in the NAcc shell, using viral-mediated RNAi techniques, reduces ethanol self-administration and preference, which is accompanied by a reduction of the reinforcing effects of ethanol (Rewal et al., 2011; Rewal et al., 2009). In addition, Nie et al. (2011) have shown that selective reduction of the GABAAR δ -subunit expression specifically in the medial NAcc shell decreases ethanol intake. Interestingly, GABAAR δ-subunits can associate with α4- and β-subunits to form a functional α4 βδ GABAAR isoform predominantly located at perisynaptic or extrasynaptic sites (Nusser, Sieghart, & Somogyi, 1998; Wei, Zhang, Peng, Houser, & Mody, 2003). The α4βδ GABAARs are highly sensitive to GABA and are responsible for the tonic inhibition in the presence of low extracellular GABA that regulates neuronal excitability (Farrant & Nusser, 2005). Low concentrations of ethanol selectively increase the tonic current mediated by δ subunit-containing GABAARs (Sundstrom-Poromaa et al., 2002; Wallner, Hanchar, & Olsen, 2003; Wei, Faria, & Mody, 2004). Tonic currents recorded in NAcc MSNs are also potentiated by low concentrations of acute ethanol (Liang, Suryanarayanan, Olsen, & Spigelman, 2009). Taken together, these results suggest that α4βδ GABAARs in the medial NAcc shell play an important role in alcohol drinking behavior. Another recent study demonstrated that inactivation of the NAcc shell by infusion of muscimol produces differential effects on the intake of ethanol, sucrose or ethanol/sucrose solutions (Stratford &Wirtshafter, 2011). This finding confirms the importance of the NAcc shell in mediating voluntary ethanol intake. All together, these studies show that the alterations of GABAergic receptors producedin the NAcc shellin responseto alcohol are determinant in the reinforcing properties of alcohol.

Chronic alcohol effects on integration of reward information processing in mesolimbic circuits

Alcohol-induced modulation of neuronal intrinsic membrane properties and synaptic transmission in NAcc microcircuits could alter the occurrence of suprathreshold postsynaptic depolarizations triggering APs. The AP constitutes the pertinent physiological coding event required for the induction of spike-time dependent plasticity (STDP), an activity-dependent bi-directional long-term synaptic plasticity, that can modulate the strength of synaptic inputs in MSNs and interneurons (Fino, Deniau, & Venance, 2008, 2009; Fino, Glowinski, & Venance, 2005; Fino & Venance, 2010; Pawlak & Kerr, 2008; Shen, Flajolet, Greengard, & Surmeier, 2008). The potentiation of glutamatergic inputs could enhance the integration of dendritic information and increase the likelihood of AP initiation at the soma (Xu, Ye, Poo, & Zhang, 2006). Alcohol-induced increase in excitatory postsynaptic inputs could promote the occurrence and the magnitude of STDP in the NAcc neurons. In contrast, inhibitory GABAergic inputs can modulate the degree of depolarizations elicited by excitatory inputs (Fitzpatrick, Akopian, & Walsh, 2001), thereby influencing the firing of MSNs and the occurrence of STDP. The modulation of the NAcc neuronal network plasticity by alcohol could lead to important alterations in the integration and transmission of pertinent information to output structures, such as the VTA which plays a major role in the behavioral effects of alcohol (Hyman et al., 2006; Nestler, 2001).

It is essential to take into account the possible effects of ethanol on the electrical properties and synaptic transmission of NAcc interneurons since these interneurons are directly involved in the regulation of NAcc microcircuits and control information processing of MSNs. Although cholinergic interneurons constitute less than 1% of the local neural population within the NAcc (Rymar, Sasseville, Luk, & Sadikot, 2004), a recent study has demonstrated their importance in reward-related responses to cocaine (Witten et al., 2010). Optogenetic inhibition of cholinergic interneurons increases MSN electrical activity in the NAcc and suppresses cocaine conditioning in freely moving rats (Witten et al., 2010). Such studies reinforce the importance of NAcc microcircuits in mediating drug-related behaviors.

In vivo studies have shown that alcohol modulates the release of dopamine in the NAcc. Acute ethanol self-administration increases the activity of VTA dopaminergic neurons producing an increase in dopamine release within the NAcc involved in the rewarding effect of ethanol (Brodie, Shefner, & Dunwiddie, 1990; Melendez et al., 2002; Weiss, Lorang, Bloom, & Koob, 1993; Yim & Gonzales, 2000). In contrast to acute ethanol, withdrawal from repeated ethanol administration is characterized by decreased levels of dopamine in the NAcc associated with a decrease in the activity of VTA dopaminergic neurons (Bailey, O'Callaghan, Croft, Manley, & Little, 2001; Diana, Pistis, Carboni, Gessa, & Rosseti, 1993; Shen, 2003; Shen & Chiodo, 1993). This depletion of dopamine in the mesolimbic system is believed to be responsible, at least in part, for dysphoria and anhedonia associated with alcohol protracted withdrawal which may trigger the drug-seeking behavior characteristic of addicts (Weiss & Porrino, 2002). Indeed, it has been shown that alcohol-dependent rats are more motivated during withdrawal to obtain ethanol in an operant self-administration task than non-dependents, and that ethanol consumption reverses the withdrawal-associated decreased dopamine levels in the NAcc (Weiss et al., 1996). The reciprocal connections between the NAcc and the VTA are critical in the regulation of dopamine release (Rahman & McBride, 2000, 2002). Thus, decreases in dopamine levels during withdrawal may result at least in part from the long-term effects of ethanol on the NAcc microcircuitry presented above. Further, ethanol-induced modifications of the VTA microcircuitry will likely affect both dopaminergic and GABAergic neurotrans-mission in the NAcc (Hopf et al., 2007; Hyman et al., 2006; Ortiz et al., 1995; Stuber et al., 2008). Similarly, excitatory inputs into the NAcc from the prelimbic cortex, hippocampus and amygdala will differentially affect reward information processing in the NAcc. Therefore, it is essential to consider the multiple effects of alcohol on the different brain structures in order to have a better understanding of the mechanisms responsible for the alcohol-induced neuroadaptations in the mesolimbic dopamine system.

While it may be tempting to speculate on what the demonstrated alcohol-induced alterations in membrane properties and synaptic transmission of MSNs might do to affect reward information processing in the NAcc circuitry, such speculation is premature. Hypothesis-driven experiments which utilize selective activation (or suppression) of various neuronal pathways of mesolimbic circuitry should be able to provide answers to these questions in the near future.

Conclusions

Collectively, the results from the studies described above provide compelling evidence that MSNs in the NAcc play a key role in mediating behavioral effects of alcohol. These studies have demonstrated that acute and chronic ethanol exposure alters electrical membrane properties of MSNs, and produces a dysregulation of glutamatergic and GABAergic transmission in the NAcc via intricate pre- and postsynaptic mechanisms. In addition, several studies have been able to relate the neuroadaptive changes produced by ethanol in the NAcc to several important behavioral symptoms of alcohol dependence such as alcohol consumption and alcohol-seeking behaviors.

Although the mechanisms involved in short-term effects of alcohol have been largely described, the long-lasting neuroadaptive changes induced by alcohol in the NAcc during protracted withdrawal from chronic alcohol exposure are still unclear. These long-lasting neuroadaptations are more likely to contribute to the development of the allostatic state mediating susceptibility to cravings and relapse. A better understanding of alcohol-induced long-lasting neuroadaptive changes in the different sub-regions of the NAcc should be useful in the development of specific pharmacological compounds to reduce alcohol-seeking behavior during protracted withdrawal.

Acknowledgments

This work was supported by the National Institutes of Health/National Institute of Alcohol Abuse and Alcoholism grants AA016100 and AA017581.

References

  1. Adams CL, Short JL, Lawrence AJ. Cue-conditioned alcohol seeking in rats following abstinence: involvement of metabotropic glutamate 5 receptors. British Journal of Pharmacology. 2010;159:534–542. doi: 10.1111/j.1476-5381.2009.00562.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Addolorato G, Leggio L, Abenavoli L, Agabio R, Caputo F, Capristo E, et al. Baclofen in the treatment of alcohol withdrawal syndrome: a comparative study vs diazepam. American Journal of Medicine. 2006;119(276):e213–278. doi: 10.1016/j.amjmed.2005.08.042. [DOI] [PubMed] [Google Scholar]
  3. Backstrom P, Hyytia P. Ionotropic glutamate receptor antagonists modulate cue-induced reinstatement of ethanol-seeking behavior. Alcoholism: Clinical & Experimental Research. 2004;28:558–565. doi: 10.1097/01.alc.0000122101.13164.21. [DOI] [PubMed] [Google Scholar]
  4. Backstrom P, Hyytia P. Suppression of alcohol self-administration and cue-induced reinstatement of alcohol seeking by the mGlu2/3 receptor agonist LY379268 and the mGlu8 receptor agonist (S)-3,4-DCPG. European Journal of Pharmacology. 2005;528:110–118. doi: 10.1016/j.ejphar.2005.10.051. [DOI] [PubMed] [Google Scholar]
  5. Bailey CP, Molleman A, Little HJ. Comparison of the effects of drugs on hyperexcitability induced in hippocampal slices by withdrawal from chronic ethanol consumption. British Journal of Pharmacology. 1998;123:215–222. doi: 10.1038/sj.bjp.0701596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bailey CP, O'Callaghan MJ, Croft AP, Manley SJ, Little HJ. Alterations in mesolimbic dopamine function during the abstinence period following chronic ethanol consumption. Neuropharmacology. 2001;41:989–999. doi: 10.1016/s0028-3908(01)00146-0. [DOI] [PubMed] [Google Scholar]
  7. Baker DA, McFarland K, Lake RW, Shen H, Tang XC, Toda S, et al. Neuroadaptations in cystine-glutamate exchange underlie cocaine relapse. Nature Neuroscience. 2003;6:743–749. doi: 10.1038/nn1069. [DOI] [PubMed] [Google Scholar]
  8. Baker DA, Xi ZX, Shen H, Swanson CJ, Kalivas PW. The origin and neuronal function of in vivo nonsynaptic glutamate. Journal of Neuroscience. 2002;22:9134–9141. doi: 10.1523/JNEUROSCI.22-20-09134.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Baptista MA, Martin-Fardon R, Weiss F. Preferential effects of the metabotropic glutamate 2/3 receptor agonist LY379268 on conditioned reinstatement versus primary reinforcement: comparison between cocaine and a potent conventional reinforcer. Journal of Neuroscience. 2004;24:4723–4727. doi: 10.1523/JNEUROSCI.0176-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Beckstead RM, Domesick VB, Nauta WJ. Efferent connections of the substantia nigra and ventral tegmental area in the rat. Brain Research. 1979;175:191–217. doi: 10.1016/0006-8993(79)91001-1. [DOI] [PubMed] [Google Scholar]
  11. Belleau ML, Warren RA. Postnatal development of electrophysiological properties of nucleus accumbens neurons. Jounal of Neurophysiology. 2000;84:2204–2216. doi: 10.1152/jn.2000.84.5.2204. [DOI] [PubMed] [Google Scholar]
  12. Bennett BD, Bolam JP. Synaptic input and output of parvalbumin-immunoreactive neurons in the neostriatum of the rat. Neuroscience. 1994;62:707–719. doi: 10.1016/0306-4522(94)90471-5. [DOI] [PubMed] [Google Scholar]
  13. Bernier BE, Whitaker LR, Morikawa H. Previous ethanol experience enhances synaptic plasticity of NMDA receptors in the ventral tegmental area. Journal of Neuroscience. 2011;31:5205–5212. doi: 10.1523/JNEUROSCI.5282-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Berton F, Francesconi WG, Madamba SG, Zieglgansberger W, Siggins GR. Acamprosate enhances N-methyl-D-apartate receptor-mediated neurotransmission but inhibits presynaptic GABA(B) receptors in nucleus accumbens neurons. Alcoholism: Clinical & Experimental Research. 1998;22:183–191. [PubMed] [Google Scholar]
  15. Blednov YA, Harris RA. Metabotropic glutamate receptor 5 (mGluR5) regulation of ethanol sedation, dependence and consumption: relationship to acamprosate actions. International Journal of Neuropsychopharmacology. 2008;11:775–793. doi: 10.1017/S1461145708008584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Blomeley CP, Cains S, Smith R, Bracci E. Ethanol affects striatal interneurons directly and projection neurons through a reduction in cholinergic tone. Neuropsychopharmacology. 2011;36:1033–1046. doi: 10.1038/npp.2010.241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Breese GR, Chu K, Dayas CV, Funk D, Knapp DJ, Koob GF, et al. Stress enhancement of craving during sobriety: a risk for relapse. Alcoholism: Clinical & Experimental Research. 2005;29:185–195. doi: 10.1097/01.alc.0000153544.83656.3c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Brodie MS, McElvain MA, Bunney EB, Appel SB. Pharmacological reduction of small conductance calcium-activated potassium current (SK) potentiates the excitatory effect of ethanol on ventral tegmental area dopamine neurons. Journal of Pharmacology and Experimental Therapeutics. 1999;290:325–333. [PubMed] [Google Scholar]
  19. Brodie MS, Scholz A, Weiger TM, Dopico AM. Ethanol interactions with calcium-dependent potassium channels. Alcoholism: Clinical & Experimental Research. 2007;31:1625–1632. doi: 10.1111/j.1530-0277.2007.00469.x. [DOI] [PubMed] [Google Scholar]
  20. Brodie MS, Shefner SA, Dunwiddie TV. Ethanol increases the firing rate of dopamine neurons of the rat ventral tegmental area in vitro. Brain Research. 1990;508:65–69. doi: 10.1016/0006-8993(90)91118-z. [DOI] [PubMed] [Google Scholar]
  21. Brog JS, Salyapongse A, Deutch AY, Zahm DS. The patterns of afferent innervation of the core and shell in the “accumbens” part of the rat ventral striatum: immunohistochemical detection of retrogradely transported fluoro-gold. Journal of Comparative Neurology. 1993;338:255–278. doi: 10.1002/cne.903380209. [DOI] [PubMed] [Google Scholar]
  22. Calton JL, Wilson WA, Moore SD. Reduction of voltage-dependent currents by ethanol contributes to inhibition of NMDA receptor-mediated excitatory synaptic transmission. Brain Research. 1999;816:142–148. doi: 10.1016/s0006-8993(98)01144-5. [DOI] [PubMed] [Google Scholar]
  23. Cao Y, Dreixler JC, Roizen JD, Roberts MT, Houamed KM. Modulation of recombinant small-conductance Ca(2+)-activated K(+) channels by the muscle relaxant chlorzoxazone and structurally related compounds. Journal of Pharmacology and Experimental Therapeutics. 2001;296:683–689. [PubMed] [Google Scholar]
  24. Carlen PL, Jahromi SS, Valiante TA, Velumian AA, Weiner JL, Zhang L. The role of calcium homeostasis and calcium currents in ethanol actions on central mammalian neurons. Alcohol and Alcoholism. 1993;2(Suppl. 2):395–401. [PubMed] [Google Scholar]
  25. Chang HT, Kitai ST. Projection neurons of the nucleus accumbens: an intracellular labeling study. Brain Research. 1985;347:112–116. doi: 10.1016/0006-8993(85)90894-7. [DOI] [PubMed] [Google Scholar]
  26. Chaudhri N, Sahuque LL, Cone JJ, Janak PH. Reinstated ethanol-seeking in rats is modulated by environmental context and requires the nucleus accumbens core. European Journal of Neuroscience. 2008;28:2288–2298. doi: 10.1111/j.1460-9568.2008.06517.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Chaudhri N, Sahuque LL, Schairer WW, Janak PH. Separable roles of the nucleus accumbens core and shell in context- and cue-induced alcohol-seeking. Neuropsychopharmacology. 2010;35:783–791. doi: 10.1038/npp.2009.187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Conrad KL, Tseng KY, Uejima JL, Reimers JM, Heng LJ, Shaham Y, et al. Formation of accumbens GluR2-lacking AMPA receptors mediates incubation of cocaine craving. Nature. 2008;454:118–121. doi: 10.1038/nature06995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Criado JR, Lee RS, Berg GI, Henriksen SJ. Sensitivity of nucleus accumbens neurons in vivo to intoxicating doses of ethanol. Alcoholism: Clinical & Experimental Research. 1995;19:164–169. doi: 10.1111/j.1530-0277.1995.tb01486.x. [DOI] [PubMed] [Google Scholar]
  30. Criado JR, Lee RS, Berg GI, Henriksen SJ. Ethanol inhibits single-unit responses in the nucleus accumbens evoked by stimulation of the basolateral nucleus of the amygdala. Alcoholism: Clinical & Experimental Research. 1997;21:368–374. [PubMed] [Google Scholar]
  31. Dahchour A, Hoffman A, Deitrich R, de Witte P. Effects of ethanol on extracellular amino acid levels in high-and low-alcohol sensitive rats: a microdialysis study. Alcohol and Alcoholism. 2000;35:548–553. doi: 10.1093/alcalc/35.6.548. [DOI] [PubMed] [Google Scholar]
  32. Dahlstrom A, Fuxe K. Evidence for the existence of an outflow of noradrenaline nerve fibres in the ventral roots of the rat spinal cord. Experientia. 1965;21:409–410. doi: 10.1007/BF02139774. [DOI] [PubMed] [Google Scholar]
  33. Danbolt NC. Glutamate uptake. Progress in Neurobiology. 2001;65:1–105. doi: 10.1016/s0301-0082(00)00067-8. [DOI] [PubMed] [Google Scholar]
  34. David HN, Ansseau M, Abraini JH. Dopamine-glutamate reciprocal modulation of release and motor responses in the rat caudate-putamen and nucleus accumbens of “intact” animals. Brain Research Reviews. 2005;50:336–360. doi: 10.1016/j.brainresrev.2005.09.002. [DOI] [PubMed] [Google Scholar]
  35. Davies AG, Pierce-Shimomura JT, Kim H, VanHoven MK, Thiele TR, Bonci A, et al. A central role of the BK potassium channel in behavioral responses to ethanol in C. elegans. Cell. 2003;115:655–666. doi: 10.1016/s0092-8674(03)00979-6. [DOI] [PubMed] [Google Scholar]
  36. de Quidt ME, Emson PC. Distribution of neuropeptide Y-like immunoreactivity in the rat central nervous systeme–II. Immunohistochemical analysis. Neuroscience. 1986;18:545–618. doi: 10.1016/0306-4522(86)90057-6. [DOI] [PubMed] [Google Scholar]
  37. Deng C, Li KY, Zhou C, Ye JH. Ethanol enhances glutamate transmission by retrograde dopamine signaling in a postsynaptic neuron/synaptic bouton preparation from the ventral tegmental area. Neuropsychopharmacology. 2009;34:1233–1244. doi: 10.1038/npp.2008.143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Diana M, Pistis M, Carboni S, Gessa GL, Rossetti ZL. Profound decrement of mesolimbic dopaminergic neuronal activity during ethanol withdrawal syndrome in rats: electrophysiological and biochemical evidence. Proceedings of the National Academy of Sciences of the United States of America. 1993;90:7966–7969. doi: 10.1073/pnas.90.17.7966. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Dolin S, Little H, Hudspith M, Pagonis C, Littleton J. Increased dihydropyridine-sensitive calcium channels in rat brain may underlie ethanol physical dependence. Neuropharmacology. 1987;26:275–279. doi: 10.1016/0028-3908(87)90220-6. [DOI] [PubMed] [Google Scholar]
  40. Dopico AM, Lemos JR, Treistman SN. Ethanol increases the activity of large conductance, Ca(2+)-activated K+ channels in isolated neurohypophysial terminals. Molecular Pharmacology. 1996;49:40–48. [PubMed] [Google Scholar]
  41. Dreixler JC, Jenkins A, Cao YJ, Roizen JD, Houamed KM. Patch-clamp analysis of anesthetic interactions with recombinant SK2 subtype neuronal calcium-activated potassium channels. Anesthesia and Analgesia. 2000;90:727–732. doi: 10.1097/00000539-200003000-00040. [DOI] [PubMed] [Google Scholar]
  42. Fallon JH, Moore RY. Catecholamine innervation of the basal forebrain. IV. Topography of the dopamine projection to the basal forebrain and neostriatum. Journal of Comparative Neurology. 1978;180:545–580. doi: 10.1002/cne.901800310. [DOI] [PubMed] [Google Scholar]
  43. Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABA(A) receptors. Nature Reviews Neuroscience. 2005;6:215–229. doi: 10.1038/nrn1625. [DOI] [PubMed] [Google Scholar]
  44. Fino E, Deniau JM, Venance L. Cell-specific spike-timing-dependent plasticity in GABAergic and cholinergic interneurons in corticostriatal rat brain slices. Journal of Physiology. 2008;586:265–282. doi: 10.1113/jphysiol.2007.144501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Fino E, Deniau JM, Venance L. Brief subthreshold events can act as Hebbian signals for long-term plasticity. PLoS One. 2009;4:e6557. doi: 10.1371/journal.pone.0006557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Fino E, Glowinski J, Venance L. Bidirectional activity-dependent plasticity at corticostriatal synapses. Journal of Neuroscience. 2005;25:11279–11287. doi: 10.1523/JNEUROSCI.4476-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Fino E, Venance L. Spike-timing dependent plasticity in striatal interneurons. Neuropharmacology. 2010;60:780–788. doi: 10.1016/j.neuropharm.2011.01.023. [DOI] [PubMed] [Google Scholar]
  48. Fitzpatrick JS, Akopian G, Walsh JP. Short-term plasticity at inhibitory synapses in rat striatum and its effects on striatal output. Journal of Neurophysiology. 2001;85:2088–2099. doi: 10.1152/jn.2001.85.5.2088. [DOI] [PubMed] [Google Scholar]
  49. Follesa P, Ticku MK. Chronic ethanol treatment differentially regulates NMDA receptor subunit mRNA expression in rat brain. Molecular Brain Research. 1995;29:99–106. doi: 10.1016/0169-328x(94)00235-7. [DOI] [PubMed] [Google Scholar]
  50. Fuchs RA, Evans KA, Parker MC, See RE. Differential involvement of the core and shell subregions of the nucleus accumbens in conditioned cue-induced reinstatement of cocaine seeking in rats. Psychopharmacology (Berlin) 2004;176:459–465. doi: 10.1007/s00213-004-1895-6. [DOI] [PubMed] [Google Scholar]
  51. Fuchs RA, Ramirez DR, Bell GH. Nucleus accumbens shell and core involvement in drug context-induced reinstatement of cocaine seeking in rats. Psychopharmacology (Berlin) 2008;200:545–556. doi: 10.1007/s00213-008-1234-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Fujii S, Yamazaki Y, Sugihara T, Wakabayashi I. Acute and chronic ethanol exposure differentially affect induction of hippocampal LTP. Brain Research. 2008;1211:13–21. doi: 10.1016/j.brainres.2008.02.052. [DOI] [PubMed] [Google Scholar]
  53. Gass JT, Olive MF. Glutamatergic substrates of drug addiction and alcoholism. Biochemical Pharmacology. 2008;75:218–265. doi: 10.1016/j.bcp.2007.06.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Gass JT, Sinclair CM, Cleva RM, Widholm JJ, Olive MF. Alcohol-seeking behavior is associated with increased glutamate transmission in basolateral amygdala and nucleus accumbens as measured by glutamate-oxidase-coated biosensors. Addiction Biology. 2011;16:215–228. doi: 10.1111/j.1369-1600.2010.00262.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Grieco TM, Malhotra JD, Chen C, Isom LL, Raman IM. Open-channel block by the cytoplasmic tail of sodium channel beta4 as a mechanism for resurgent sodium current. Neuron. 2005;45:233–244. doi: 10.1016/j.neuron.2004.12.035. [DOI] [PubMed] [Google Scholar]
  56. Gu N, Vervaeke K, Storm JF. BK potassium channels facilitate high-frequency firing and cause early spike frequency adaptation in rat CA1 hippocampal pyramidal cells. Journal of Physiology. 2007;580:859–882. doi: 10.1113/jphysiol.2006.126367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Guire ES, Oh MC, Soderling TR, Derkach VA. Recruitment of calcium-permeable AMPA receptors during synaptic potentiation is regulated by CaM-kinase I. Journal of Neuroscience. 2008;28:6000–6009. doi: 10.1523/JNEUROSCI.0384-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Guppy LJ, Crabbe JC, Littleton JM. Time course and genetic variation in the regulation of calcium channel antagonist binding sites in rodent tissues during the induction of ethanol physical dependence and withdrawal. Alcohol and Alcoholism. 1995;30:607–615. [PubMed] [Google Scholar]
  59. Guppy LJ, Littleton JM. Binding characteristics of the calcium channel antagonist [3H]-nitrendipine in tissues from ethanol-dependent rats. Alcohol and Alcoholism. 1994;29:283–293. [PubMed] [Google Scholar]
  60. Hendricson AW, Maldve RE, Salinas AG, Theile JW, Zhang TA, Diaz LM, et al. Aberrant synaptic activation of N-methyl-D-aspartate receptors underlies ethanol withdrawal hyperexcitability. Journal of Pharmacology and Experimental Therapeutics. 2007;321:60–72. doi: 10.1124/jpet.106.111419. [DOI] [PubMed] [Google Scholar]
  61. Hendricson AW, Thomas MP, Lippmann MJ, Morrisett RA. Suppression of L-type voltage-gated calcium channel-dependent synaptic plasticity by ethanol: analysis of miniature synaptic currents and dendritic calcium transients. Journal of Pharmacology and Experimental Therapeutics. 2003;307:550–558. doi: 10.1124/jpet.103.055137. [DOI] [PubMed] [Google Scholar]
  62. Hoeffer CA, Klann E. mTOR signaling: at the crossroads of plasticity, memory and disease. Trends in Neuroscience. 2010;33:67–75. doi: 10.1016/j.tins.2009.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Hopf FW, Bowers MS, Chang SJ, Chen BT, Martin M, Seif T, et al. Reduced nucleus accumbens SK channel activity enhances alcohol seeking during abstinence. Neuron. 2010;65:682–694. doi: 10.1016/j.neuron.2010.02.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Hopf FW, Martin M, Chen BT, Bowers MS, Mohamedi MM, Bonci A. Withdrawal from intermittent ethanol exposure increases probability of burst firing in VTA neurons in vitro. Journal of Neurophysiology. 2007;98:2297–2310. doi: 10.1152/jn.00824.2007. [DOI] [PubMed] [Google Scholar]
  65. Hopf FW, Simms JA, Chang SJ, Seif T, Bartlett SE, Bonci A. Chlorzoxazone, an SK-type potassium channel activator used in humans, reduces excessive alcohol intake in rats. Biological Psychiatry. 2011;69:618–624. doi: 10.1016/j.biopsych.2010.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Horishita T, Harris RA. n-Alcohols inhibit voltage-gated Na+ channels expressed in Xenopus oocytes. Journal of Pharmacology and Experimental Therapeutics. 2008;326:270–277. doi: 10.1124/jpet.108.138370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Hussain Z, Johnson LR, Totterdell S. A light and electron microscopic study of NADPH-diaphorase-, calretinin- and parvalbumin-containing neurons in the rat nucleus accumbens. Journal of Chemical Neuroanatomy. 1996;10:19–39. doi: 10.1016/0891-0618(95)00098-4. [DOI] [PubMed] [Google Scholar]
  68. Hwang BH, Lumeng L, Wu JY, Li TK. Increased number of GABAergic terminals in the nucleus accumbens is associated with alcohol preference in rats. Alcoholism: Clinical & Experimental Research. 1990;14:503–507. doi: 10.1111/j.1530-0277.1990.tb01188.x. [DOI] [PubMed] [Google Scholar]
  69. Hyman SE, Malenka RC, Nestler EJ. Neural mechanisms of addiction: the role of reward-related learning and memory. Annual Review of Neuroscience. 2006;29:565–598. doi: 10.1146/annurev.neuro.29.051605.113009. [DOI] [PubMed] [Google Scholar]
  70. Ishikawa M, Mu P, Moyer JT, Wolf JA, Quock RM, Davies NM, et al. Homeostatic synapse-driven membrane plasticity in nucleus accumbens neurons. Journal of Neuroscience. 2009;29:5820–5831. doi: 10.1523/JNEUROSCI.5703-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Izumi Y, Nagashima K, Murayama K, Zorumski CF. Acute effects of ethanol on hippocampal long-term potentiation and long-term depression are mediated by different mechanisms. Neuroscience. 2005;136:509–517. doi: 10.1016/j.neuroscience.2005.08.002. [DOI] [PubMed] [Google Scholar]
  72. Jeanes ZM, Buske TR, Morrisett RA. In vivo chronic intermittent ethanol exposure reverses the polarity of synaptic plasticity in the nucleus accumbens shell. Journal of Pharmacology and Experimental Therapeutics. 2011;336:155–164. doi: 10.1124/jpet.110.171009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Johnson BA, Rosenthal N, Capece JA, Wiegand F, Mao L, Beyers K, et al. Topiramate for treating alcohol dependence: a randomized controlled trial. Journal of American Medical Association. 2007;298:1641–1651. doi: 10.1001/jama.298.14.1641. [DOI] [PubMed] [Google Scholar]
  74. Kalivas PW. The glutamate homeostasis hypothesis of addiction. Nature Reviews Neuroscience. 2009;10:561–572. doi: 10.1038/nrn2515. [DOI] [PubMed] [Google Scholar]
  75. Kalivas PW, Churchill L, Klitenick MA. GABA and enkephalin projection from the nucleus accumbens and ventral pallidum to the ventral tegmental area. Neuroscience. 1993;57:1047–1060. doi: 10.1016/0306-4522(93)90048-k. [DOI] [PubMed] [Google Scholar]
  76. Kalivas PW, Lalumiere RT, Knackstedt L, Shen H. Glutamate transmission in addiction. Neuropharmacology. 2009;56(Suppl. 1):169–173. 3. doi: 10.1016/j.neuropharm.2008.07.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Kapasova Z, Szumlinski KK. Strain differences in alcohol-induced neurochemical plasticity: a role for accumbens glutamate in alcohol intake. Alcoholism: Clinical & Experimental Research. 2008;32:617–631. doi: 10.1111/j.1530-0277.2008.00620.x. [DOI] [PubMed] [Google Scholar]
  78. Kim J, Park BH, Lee JH, Park SK, Kim JH. Cell type-specific alterations in the nucleus accumbens by repeated exposures to cocaine. Biological Psychiatry. 2011;69:1026–1034. doi: 10.1016/j.biopsych.2011.01.013. [DOI] [PubMed] [Google Scholar]
  79. Knackstedt LA, Melendez RI, Kalivas PW. Ceftriaxone restores glutamate homeostasis and prevents relapse to cocaine seeking. Biological Psychiatry. 2010;67:81–84. doi: 10.1016/j.biopsych.2009.07.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Koob GF. Alcoholism: allostasis and beyond. Alcoholism: Clinical & Experimental Research. 2003;27:232–243. doi: 10.1097/01.ALC.0000057122.36127.C2. [DOI] [PubMed] [Google Scholar]
  81. Koob GF, Nestler EJ. The neurobiology of drug addiction. Journal of Neuropsychiatry & Clinical Neurosciences. 1997;9:482–497. doi: 10.1176/jnp.9.3.482. [DOI] [PubMed] [Google Scholar]
  82. Kumar S, Fleming RL, Morrow AL. Ethanol regulation of gamma-aminobutyric acid A receptors: genomic and nongenomic mechanisms. Pharmacology & Therapeutics. 2004;101:211–226. doi: 10.1016/j.pharmthera.2003.12.001. [DOI] [PubMed] [Google Scholar]
  83. Kumar S, Porcu P, Werner DF, Matthews DB, Diaz-Granados JL, Helfand RS, et al. The role of GABA(A) receptors in the acute and chronic effects of ethanol: a decade of progress. Psychopharmacology (Berlin) 2009;205:529–564. doi: 10.1007/s00213-009-1562-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Kumari M, Ticku MK. Regulation of NMDA receptors by ethanol. Progress in Drug Research. 2000;54:152–189. [PubMed] [Google Scholar]
  85. Lack AK, Diaz MR, Chappell A, DuBois DW, McCool BA. Chronic ethanol and withdrawal differentially modulate pre- and postsynaptic function at glutamatergic synapses in rat basolateral amygdala. Journal of Neurophysiology. 2007;98:3185–3196. doi: 10.1152/jn.00189.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Lallemand F, Ward RJ, De Witte P, Verbanck P. Binge drinking +/− chronic nicotine administration alters extracellular glutamate and arginine levels in the nucleus accumbens of adult male and female wistar rats. Alcohol and Alcoholism. 2011;46:373–382. doi: 10.1093/alcalc/agr031. [DOI] [PubMed] [Google Scholar]
  87. Le Moal M. Drug abuse: vulnerability and transition to addiction. Pharmacopsychiatry. 2009;42(Suppl. 1):S42–55. doi: 10.1055/s-0029-1216355. [DOI] [PubMed] [Google Scholar]
  88. Liang J, Suryanarayanan A, Abriam A, Snyder B, Olsen RW, Spigelman I. Mechanisms of reversible GABAA receptor plasticity after ethanol intoxication. Journal of Neuroscience. 2007;27:12367–12377. doi: 10.1523/JNEUROSCI.2786-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Liang J, Suryanarayanan A, Olsen RW, Spigelman I. Ethanol-induced plasticity of GABAA receptors in different brain regions in rat. Alcoholism: Clinical & Experimental Research. 2009;33(Suppl. 84A):294. [Google Scholar]
  90. Liang J, Zhang N, Cagetti E, Houser CR, Olsen RW, Spigelman I. Chronic intermittent ethanol-induced switch of ethanol actions from extrasynaptic to synaptic hippocampal GABAA receptors. Journal of Neuroscience. 2006;26:1749–1758. doi: 10.1523/JNEUROSCI.4702-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Liu J, Wang L. Baclofen for alcohol withdrawal. Cochrane Database of Systematic Reviews. 2011:008502. doi: 10.1002/14651858.CD008502.pub2. [DOI] [PubMed] [Google Scholar]
  92. Liu YC, Lo YK, Wu SN. Stimulatory effects of chlorzoxazone, a centrally acting muscle relaxant, on large conductance calcium-activated potassium channels in pituitary GH3 cells. Brain Research. 2003;959:86–97. doi: 10.1016/s0006-8993(02)03730-7. [DOI] [PubMed] [Google Scholar]
  93. Lovinger DM. Alcohols and neurotransmitter gated ion channels: past, present and future. Naunyn Schmiedebergs Archives of Pharmacology. 1997;356:267–282. doi: 10.1007/pl00005051. [DOI] [PubMed] [Google Scholar]
  94. Madayag A, Lobner D, Kau KS, Mantsch JR, Abdulhameed O, Hearing M, et al. Repeated N-acetylcysteine administration alters plasticity-dependent effects of cocaine. Journal of Neuroscience. 2007;27:13968–13976. doi: 10.1523/JNEUROSCI.2808-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Maldve RE, Zhang TA, Ferrani-Kile K, Schreiber SS, Lippmann MJ, Snyder GL, et al. DARPP-32 and regulation of the ethanol sensitivity of NMDA receptors in the nucleus accumbens. Nature Neuroscience. 2002;5:641–648. doi: 10.1038/nn877. [DOI] [PubMed] [Google Scholar]
  96. Mameli M, Balland B, Lujan R, Luscher C. Rapid synthesis and synaptic insertion of GluR2 for mGluR-LTD in the ventral tegmental area. Science. 2007;317:530–533. doi: 10.1126/science.1142365. [DOI] [PubMed] [Google Scholar]
  97. Mameli M, Halbout B, Creton C, Engblom D, Parkitna JR, Spanagel R, et al. Cocaine-evoked synaptic plasticity: persistence in the VTA triggers adaptations in the NAc. Nature Neuroscience. 2009;12:1036–1041. doi: 10.1038/nn.2367. [DOI] [PubMed] [Google Scholar]
  98. Martin G, Puig S, Pietrzykowski A, Zadek P, Emery P, Treistman S. Somatic localization of a specific large-conductance calcium-activated potassium channel subtype controls compartmentalized ethanol sensitivity in the nucleus accumbens. Journal of Neuroscience. 2004;24:6563–6572. doi: 10.1523/JNEUROSCI.0684-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Martin GE, Hendrickson LM, Penta KL, Friesen RM, Pietrzykowski AZ, Tapper AR, et al. Identification of a BK channel auxiliary protein controlling molecular and behavioral tolerance to alcohol. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:17543–17548. doi: 10.1073/pnas.0801068105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Marty VN, Spigelman I. Contrasting effects of chronic intermittent ethanol treatment on membrane properties and synaptic transmission in the nucleus accumbens and the paraventricular nucleus. Washington DC: Society for Neuroscience 2011 Online; 2011. Program No. 373.14 Abstract Viewer Itinerary Planner. [Google Scholar]
  101. Melendez RI, Hicks MP, Cagle SS, Kalivas PW. Ethanol exposure decreases glutamate uptake in the nucleus accumbens. Alcoholism: Clinical & Experimental Research. 2005;29:326–333. doi: 10.1097/01.alc.0000156086.65665.4d. [DOI] [PubMed] [Google Scholar]
  102. Melendez RI, Rodd-Henricks ZA, Engleman EA, Li TK, McBride WJ, Murphy JM. Microdialysis of dopamine in the nucleus accumbens of alcohol-preferring (P) rats during anticipation and operant self-administration of ethanol. Alcoholism: Clinical & Experimental Research. 2002;26:318–325. [PubMed] [Google Scholar]
  103. Miguens M, Del Olmo N, Higuera-Matas A, Torres I, Garcia-Lecumberri C, Ambrosio E. Glutamate and aspartate levels in the nucleus accumbens during cocaine self-administration and extinction: a time course microdialysis study. Psychopharmacology (Berlin) 2008;196:303–313. doi: 10.1007/s00213-007-0958-x. [DOI] [PubMed] [Google Scholar]
  104. Moran MM, McFarland K, Melendez RI, Kalivas PW, Seamans JK. Cystine/glutamate exchange regulates metabotropic glutamate receptor presynaptic inhibition of excitatory transmission and vulnerability to cocaine seeking. Journal of Neuroscience. 2005;25:6389–6393. doi: 10.1523/JNEUROSCI.1007-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Morrisett RA, Swartzwelder HS. Attenuation of hippocampal long-term potentiation by ethanol: a patch-clamp analysis of glutamatergic and GABAergic mechanisms. Journal of Neuroscience. 1993;13:2264–2272. doi: 10.1523/JNEUROSCI.13-05-02264.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Moussawi K, Pacchioni A, Moran M, Olive MF, Gass JT, Lavin A, et al. N-Acetylcysteine reverses cocaine-induced metaplasticity. Nature Neuroscience. 2009;12:182–189. doi: 10.1038/nn.2250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Mu P, Moyer JT, Ishikawa M, Zhang Y, Panksepp J, Sorg BA, et al. Exposure to cocaine dynamically regulates the intrinsic membrane excitability of nucleus accumbens neurons. Journal of Neuroscience. 2010;30:3689–3699. doi: 10.1523/JNEUROSCI.4063-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Mulholland PJ, Becker HC, Woodward JJ, Chandler LJ. Small conductance calcium-activated potassium type 2 channels regulate alcohol-associated plasticity of glutamatergic synapses. Biological Psychiatry. 2011;69:625–632. doi: 10.1016/j.biopsych.2010.09.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Mulholland PJ, Hopf FW, Bukiya AN, Martin GE, Liu J, Dopico AM, et al. Sizing up ethanol-induced plasticity: the role of small and large conductance calcium-activated potassium channels. Alcoholism: Clinical & Experimental Research. 2009;33:1125–1135. doi: 10.1111/j.1530-0277.2009.00936.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Mulligan MK, Ponomarev I, Hitzemann RJ, Belknap JK, Tabakoff B, Harris RA, et al. Toward understanding the genetics of alcohol drinking through transcriptome meta-analysis. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:6368–6373. doi: 10.1073/pnas.0510188103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Mulligan MK, Rhodes JS, Crabbe JC, Mayfield RD, Adron Harris R, Ponomarev I. Molecular profiles of drinking alcohol to intoxication in C57BL/6J mice. Alcoholism: Clinical & Experimental Research. 2011;35:659–670. doi: 10.1111/j.1530-0277.2010.01384.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Nauta WJ, Smith GP, Faull RL, Domesick VB. Efferent connections and nigral afferents of the nucleus accumbens septi in the rat. Neuroscience. 1978;3:385–401. doi: 10.1016/0306-4522(78)90041-6. [DOI] [PubMed] [Google Scholar]
  113. Neasta J, Ben Hamida S, Yowell Q, Carnicella S, Ron D. Role for mammalian target of rapamycin complex 1 signaling in neuroadaptations underlying alcohol-related disorders. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:20093–20098. doi: 10.1073/pnas.1005554107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Neasta J, Hamida SB, Yowell QV, Carnicella S, Ron D. AKT signaling pathway in the nucleus accumbens mediates excessive alcohol drinking behaviors. Biological Psychiatry. 2011 doi: 10.1016/j.biopsych.2011.03.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Nestler EJ. Molecular neurobiology of addiction. American Journal on Addictions. 2001;10:201–217. doi: 10.1080/105504901750532094. [DOI] [PubMed] [Google Scholar]
  116. Nie H, Rewal M, Gill TM, Ron D, Janak PH. Extrasynaptic delta-containing GABAA receptors in the nucleus accumbens dorsomedial shell contribute to alcohol intake. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:4459–4464. doi: 10.1073/pnas.1016156108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Nie Z, Madamba SG, Siggins GR. Ethanol inhibits glutamatergic neurotransmission in nucleus accumbens neurons by multiple mechanisms. Journal of Pharmacology and Experimental Therapeutics. 1994;271:1566–1573. [PubMed] [Google Scholar]
  118. Nie Z, Madamba SG, Siggins GR. Ethanol enhances gamma-aminobutyric acid responses in a subpopulation of nucleus accumbens neurons: role of metabotropic glutamate receptors. Journal of Pharmacology and Experimental Therapeutics. 2000;293:654–661. [PubMed] [Google Scholar]
  119. Nisenbaum ES, Wilson CJ. Potassium currents responsible for inward and outward rectification in rat neostriatal spiny projection neurons. Journal of Neuroscience. 1995;15:4449–4463. doi: 10.1523/JNEUROSCI.15-06-04449.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Nisenbaum ES, Xu ZC, Wilson CJ. Contribution of a slowly inactivating potassium current to the transition to firing of neostriatal spiny projection neurons. Journal of Neurophysiology. 1994;71:1174–1189. doi: 10.1152/jn.1994.71.3.1174. [DOI] [PubMed] [Google Scholar]
  121. Nusser Z, Sieghart W, Somogyi P. Segregation of different GABAA receptors to synaptic and extrasynaptic membranes of cerebellar granule cells. Journal of Neuroscience. 1998;18:1693–1703. doi: 10.1523/JNEUROSCI.18-05-01693.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. O'Donnell P, Grace AA. Physiological and morphological properties of accumbens core and shell neurons recorded in vitro. Synapse. 1993;13:135–160. doi: 10.1002/syn.890130206. [DOI] [PubMed] [Google Scholar]
  123. O'Donnell P, Grace AA. Synaptic interactions among excitatory afferents to nucleus accumbens neurons: hippocampal gating of prefrontal cortical input. Journal of Neuroscience. 1995;15:3622–3639. doi: 10.1523/JNEUROSCI.15-05-03622.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Obara I, Bell RL, Goulding SP, Reyes CM, Larson LA, Ary AW, et al. Differential effects of chronic ethanol consumption and withdrawal on homer/glutamate receptor expression in subregions of the accumbens and amygdala of P rats. Alcoholism: Clinical & Experimental Research. 2009;33:1924–1934. doi: 10.1111/j.1530-0277.2009.01030.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Ortiz J, Fitzgerald LW, Charlton M, Lane S, Trevisan L, Guitart X, et al. Biochemical actions of chronic ethanol exposure in the mesolimbic dopamine system. Synapse. 1995;21:289–298. doi: 10.1002/syn.890210403. [DOI] [PubMed] [Google Scholar]
  126. Papadeas S, Grobin AC, Morrow AL. Chronic ethanol consumption differentially alters GABA(A) receptor alpha1 and alpha4 subunit peptide expression and GABA(A) receptor-mediated 36 Cl(−) uptake in mesocorticolimbic regions of rat brain. Alcoholism: Clinical & Experimental Research. 2001;25:1270–1275. [PubMed] [Google Scholar]
  127. Pawlak V, Kerr JN. Dopamine receptor activation is required for corticostriatal spike-timing-dependent plasticity. Journal of Neuroscience. 2008;28:2435–2446. doi: 10.1523/JNEUROSCI.4402-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Pendyam S, Mohan A, Kalivas PW, Nair SS. Computational model of extracellular glutamate in the nucleus accumbens incorporates neuro-adaptations by chronic cocaine. Neuroscience. 2009;158:1266–1276. doi: 10.1016/j.neuroscience.2008.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Perez-Velazquez JL, Valiante TA, Carlen PL. Changes in calcium currents during ethanol withdrawal in a genetic mouse model. Brain Research. 1994;649:305–309. doi: 10.1016/0006-8993(94)91077-4. [DOI] [PubMed] [Google Scholar]
  130. Peters J, Kalivas PW. The group II metabotropic glutamate receptor agonist, LY379268, inhibits both cocaine- and food-seeking behavior in rats. Psychopharmacology (Berlin) 2006;186:143–149. doi: 10.1007/s00213-006-0372-9. [DOI] [PubMed] [Google Scholar]
  131. Pierce RC, Bell K, Duffy P, Kalivas PW. Repeated cocaine augments excitatory amino acid transmission in the nucleus accumbens only in rats having developed behavioral sensitization. Journal of Neuroscience. 1996;16:1550–1560. doi: 10.1523/JNEUROSCI.16-04-01550.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Pietrzykowski AZ, Friesen RM, Martin GE, Puig SI, Nowak CL, Wynne PM, et al. Posttranscriptional regulation of BK channel splice variant stability by miR-9 underlies neuroadaptation to alcohol. Neuron. 2008;59:274–287. doi: 10.1016/j.neuron.2008.05.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Ponomarev I, Maiya R, Harnett MT, Schafer GL, Ryabinin AE, Blednov YA, et al. Transcriptional signatures of cellular plasticity in mice lacking the alpha1 subunit of GABAA receptors. Journal of Neuroscience. 2006;26:5673–5683. doi: 10.1523/JNEUROSCI.0860-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Rahman S, McBride WJ. Feedback control of mesolimbic somatodendritic dopamine release in rat brain. Journal of Neurochemistry. 2000;74:684–692. doi: 10.1046/j.1471-4159.2000.740684.x. [DOI] [PubMed] [Google Scholar]
  135. Rahman S, McBride WJ. Involvement of GABA and cholinergic receptors in the nucleus accumbens on feedback control of somatodendritic dopamine release in the ventral tegmental area. Journal of Neurochemistry. 2002;80:646–654. doi: 10.1046/j.0022-3042.2001.00739.x. [DOI] [PubMed] [Google Scholar]
  136. Rewal M, Donahue R, Gill TM, Nie H, Ron D, Janak PH. Alpha4 subunit-containing GABA(A) receptors in the accumbens shell contribute to the reinforcing effects of alcohol. Addiction Biology. 2011 doi: 10.1111/j.1369-1600.2011.00333.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Rewal M, Jurd R, Gill TM, He DY, Ron D, Janak PH. Alpha4-containing GABAA receptors in the nucleus accumbens mediate moderate intake of alcohol. Journal of Neuroscience. 2009;29:543–549. doi: 10.1523/JNEUROSCI.3199-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Rossetti ZL, Isola D, De Vry J, Fadda F. Effects of nimodipine on extracellular dopamine levels in the rat nucleus accumbens in ethanol withdrawal. Neuropharmacology. 1999;38:1361–1369. doi: 10.1016/s0028-3908(99)00039-8. [DOI] [PubMed] [Google Scholar]
  139. Rymar VV, Sasseville R, Luk KC, Sadikot AF. Neurogenesis and stereological morphometry of calretinin-immunoreactive GABAergic interneu-rons of the neostriatum. Journal of Comparative Neurology. 2004;469:325–339. doi: 10.1002/cne.11008. [DOI] [PubMed] [Google Scholar]
  140. Sanchis-Segura C, Borchardt T, Vengeliene V, Zghoul T, Bachteler D, Gass P, et al. Involvement of the AMPA receptor GluR-C subunit in alcohol-seeking behavior and relapse. Journal of Neuroscience. 2006;26:1231–1238. doi: 10.1523/JNEUROSCI.4237-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Schummers J, Browning MD. Evidence for a role for GABA(A) and NMDA receptors in ethanol inhibition of long-term potentiation. Molecular Brain Research. 2001;94:9–14. doi: 10.1016/s0169-328x(01)00161-9. [DOI] [PubMed] [Google Scholar]
  142. Sesack SR, Grace AA. Cortico-Basal Ganglia reward network: micro-circuitry. Neuropsychopharmacology. 2010;35:27–47. doi: 10.1038/npp.2009.93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Shen RY. Ethanol withdrawal reduces the number of spontaneously active ventral tegmental area dopamine neurons in conscious animals. Journal of Pharmacology and Experimental Therapeutics. 2003;307:566–572. doi: 10.1124/jpet.103.053371. [DOI] [PubMed] [Google Scholar]
  144. Shen RY, Chiodo LA. Acute withdrawal after repeated ethanol treatment reduces the number of spontaneously active dopaminergic neurons in the ventral tegmental area. Brain Research. 1993;622:289–293. doi: 10.1016/0006-8993(93)90831-7. [DOI] [PubMed] [Google Scholar]
  145. Shen W, Flajolet M, Greengard P, Surmeier DJ. Dichotomous dopa-minergic control of striatal synaptic plasticity. Science. 2008;321:848–851. doi: 10.1126/science.1160575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Shiraishi M, Harris RA. Effects of alcohols and anesthetics on recombinant voltage-gated Na+ channels. Journal of Pharmacology and Experimental Therapeutics. 2004;309:987–994. doi: 10.1124/jpet.103.064063. [DOI] [PubMed] [Google Scholar]
  147. Sinha R. How does stress increase risk of drug abuse and relapse? Psychopharmacology (Berlin) 2001;158:343–359. doi: 10.1007/s002130100917. [DOI] [PubMed] [Google Scholar]
  148. Smith JW, Watson WP, Stephens DN, Little HJ. The calcium channel antagonist, nimodipine, decreases operant self-administration of low concentrations of ethanol. Behavioural Pharmacology. 1999;10:793–802. doi: 10.1097/00008877-199912000-00011. [DOI] [PubMed] [Google Scholar]
  149. Sondheimer I, Knackstedt LA. Ceftriaxone prevents the induction of cocaine sensitization and produces enduring attenuation of cue- and cocaine-primed reinstatement of cocaine-seeking. Behavioural Brain Research. 2011;225:252–258. doi: 10.1016/j.bbr.2011.07.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Steephen JE, Manchanda R. Differences in biophysical properties of nucleus accumbens medium spiny neurons emerging from inactivation of inward rectifying potassium currents. Journal of Computational Neuroscience. 2009;27:453–470. doi: 10.1007/s10827-009-0161-7. [DOI] [PubMed] [Google Scholar]
  151. Steffensen SC, Nie Z, Criado JR, Siggins GR. Ethanol inhibition of N-methyl-D-aspartate responses involves presynaptic gamma-aminobutyric acid(B) receptors. Journal of Pharmacology and Experimental Therapeutics. 2000;294:637–647. [PubMed] [Google Scholar]
  152. Stratford TR, Wirtshafter D. Opposite effects on the ingestion of ethanol and sucrose solutions after injections of muscimol into the nucleus accumbens shell. Behavioural Brain Research. 2011;216:514–518. doi: 10.1016/j.bbr.2010.08.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Stuber GD, Hopf FW, Hahn J, Cho SL, Guillory A, Bonci A. Voluntary ethanol intake enhances excitatory synaptic strength in the ventral tegmental area. Alcoholism: Clinical & Experimental Research. 2008;32:1714–1720. doi: 10.1111/j.1530-0277.2008.00749.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Stuber GD, Hopf FW, Tye KM, Chen BT, Bonci A. Neuroplastic alterations in the limbic system following cocaine or alcohol exposure. Current Topics in Behavioral Neurosciences. 2010;3:3–27. doi: 10.1007/7854_2009_23. [DOI] [PubMed] [Google Scholar]
  155. Stuber GD, Sparta DR, Stamatakis AM, van Leeuwen WA, Hardjoprajitno JE, Cho S, et al. Excitatory transmission from the amygdala to nucleus accumbens facilitates reward seeking. Nature. 2011;475:377–380. doi: 10.1038/nature10194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Sundstrom-Poromaa I, Smith DH, Gong QH, Sabado TN, Li X, Light A, et al. Hormonally regulated alpha(4)beta(2)delta GABA(A) receptors are a target for alcohol. Nature Neuroscience. 2002;5:721–722. doi: 10.1038/nn888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Swanson CJ, Baker DA, Carson D, Worley PF, Kalivas PW. Repeated cocaine administration attenuates group I metabotropic glutamate receptor-mediated glutamate release and behavioral activation: a potential role for Homer. Journal of Neuroscience. 2001;21:9043–9052. doi: 10.1523/JNEUROSCI.21-22-09043.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Swanson LW. The projections of the ventral tegmental area and adjacent regions: a combined fluorescent retrograde tracer and immunofluorescence study in the rat. Brain Research Bulletin. 1982;9:321–353. doi: 10.1016/0361-9230(82)90145-9. [DOI] [PubMed] [Google Scholar]
  159. Szumlinski KK, Ary AW, Lominac KD, Klugmann M, Kippin TE. Accumbens Homer2 overexpression facilitates alcohol-induced neuroplasticity in C57BL/6J mice. Neuropsychopharmacology. 2008;33:1365–1378. doi: 10.1038/sj.npp.1301473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Szumlinski KK, Diab ME, Friedman R, Henze LM, Lominac KD, Bowers MS. Accumbens neurochemical adaptations produced by binge-like alcohol consumption. Psychopharmacology (Berlin) 2007;190:415–431. doi: 10.1007/s00213-006-0641-7. [DOI] [PubMed] [Google Scholar]
  161. Szumlinski KK, Lominac KD, Oleson EB, Walker JK, Mason A, Dehoff MH, et al. Homer2 is necessary for EtOH-induced neuroplasticity. Journal of Neuroscience. 2005;25:7054–7061. doi: 10.1523/JNEUROSCI.1529-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Tabakoff B, Saba L, Kechris K, Hu W, Bhave SV, Finn DA, et al. The genomic determinants of alcohol preference in mice. Mammalian Genome. 2008;19:352–365. doi: 10.1007/s00335-008-9115-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Taverna S, Canciani B, Pennartz CM. Membrane properties and synaptic connectivity of fast-spiking interneurons in rat ventral striatum. Brain Research. 2007;1152:49–56. doi: 10.1016/j.brainres.2007.03.053. [DOI] [PubMed] [Google Scholar]
  164. Taverna S, van Dongen YC, Groenewegen HJ, Pennartz CM. Direct physiological evidence for synaptic connectivity between medium-sized spiny neurons in rat nucleus accumbens in situ. Journal of Neurophysiology. 2004;91:1111–1121. doi: 10.1152/jn.00892.2003. [DOI] [PubMed] [Google Scholar]
  165. Tokuda K, Izumi Y, Zorumski CF. Ethanol enhances neurosteroidogenesis in hippocampal pyramidal neurons by paradoxical NMDA receptor activation. Journal of Neuroscience. 2011;31:9905–9909. doi: 10.1523/JNEUROSCI.1660-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Treistman SN, Martin GE. BK Channels: mediators and models for alcohol tolerance. Trends in Neurosciences. 2009;32:629–637. doi: 10.1016/j.tins.2009.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Usuda I, Tanaka K, Chiba T. Efferent projections of the nucleus accumbens in the rat with special reference to subdivision of the nucleus: biotinylated dextran amine study. Brain Research. 1998;797:73–93. doi: 10.1016/s0006-8993(98)00359-x. [DOI] [PubMed] [Google Scholar]
  168. Van Bockstaele EJ, Pickel VM. GABA-containing neurons in the ventral tegmental area project to the nucleus accumbens in rat brain. Brain Research. 1995;682:215–221. doi: 10.1016/0006-8993(95)00334-m. [DOI] [PubMed] [Google Scholar]
  169. Veatch LM, Gonzalez LP. Nifedipine alleviates alterations in hippocampal kindling after repeated ethanol withdrawal. Alcoholism: Clinical & Experimental Research. 2000;24:484–491. [PubMed] [Google Scholar]
  170. Velazquez-Marrero C, Wynne P, Bernardo A, Palacio S, Martin G, Treistman SN. The relationship between duration of initial alcohol exposure and persistence of molecular tolerance is markedly nonlinear. Journal of Neuroscience. 2011;31:2436–2446. doi: 10.1523/JNEUROSCI.5429-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Vincent SR, Johansson O. Striatal neurons containing both somatostatin- and avian pancreatic polypeptide (APP)-like immunoreactivities and NADPH-diaphorase activity: a light and electron microscopic study. Journal of Comparative Neurology. 1983;217:264–270. doi: 10.1002/cne.902170304. [DOI] [PubMed] [Google Scholar]
  172. Vincent SR, Johansson O, Hokfelt T, Skirboll L, Elde RP, Terenius L, et al. NADPH-diaphorase: a selective histochemical marker for striatal neurons containing both somatostatin- and avian pancreatic polypeptide (APP)-like immunoreactivities. Journal of Comparative Neurology. 1983;217:252–263. doi: 10.1002/cne.902170303. [DOI] [PubMed] [Google Scholar]
  173. Voorn P, Jorritsma-Byham B, Van Dijk C, Buijs RM. The dopaminergic innervation of the ventral striatum in the rat: a light- and electron-microscopical study with antibodies against dopamine. Journal of Comparative Neurology. 1986;251:84–99. doi: 10.1002/cne.902510106. [DOI] [PubMed] [Google Scholar]
  174. Wallner M, Hanchar HJ, Olsen RW. Ethanol enhances alpha 4 beta 3 delta and alpha 6 beta 3 delta gamma-aminobutyric acid type A receptors at low concentrations known to affect humans. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:15218–15223. doi: 10.1073/pnas.2435171100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Walter HJ, Messing RO. Regulation of neuronal voltage-gated calcium channels by ethanol. Neurochemistry International. 1999;35:95–101. doi: 10.1016/s0197-0186(99)00050-9. [DOI] [PubMed] [Google Scholar]
  176. Wang J, Carnicella S, Phamluong K, Jeanblanc J, Ronesi JA, Chaudhri N, et al. Ethanol induces long-term facilitation of NR2B-NMDA receptor activity in the dorsal striatum: implications for alcohol drinking behavior. Journal of Neuroscience. 2007;27:3593–3602. doi: 10.1523/JNEUROSCI.4749-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Wang J, Lanfranco MF, Gibb SL, Ron D. Ethanol-mediated long-lasting adaptations of the NR2B-containing NMDA receptors in the dorsomedial striatum. Channels (Austin) 2011;5:205–209. doi: 10.4161/chan.5.3.14856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Watson WP, Little HJ. Selectivity of the protective effects of dihydropyridine calcium channel antagonists against the ethanol withdrawal syndrome. Brain Research. 2002;930:111–122. doi: 10.1016/s0006-8993(02)02236-9. [DOI] [PubMed] [Google Scholar]
  179. Wei W, Faria LC, Mody I. Low ethanol concentrations selectively augment the tonic inhibition mediated by delta subunit-containing GABAA receptors in hippocampal neurons. Journal of Neuroscience. 2004;24:8379–8382. doi: 10.1523/JNEUROSCI.2040-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Wei W, Zhang N, Peng Z, Houser CR, Mody I. Perisynaptic localization of delta subunit-containing GABA(A) receptors and their activation by GABA spillover in the mouse dentate gyrus. Journal of Neuroscience. 2003;23:10650–10661. doi: 10.1523/JNEUROSCI.23-33-10650.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Weiner JL, Valenzuela CF. Ethanol modulation of GABAergic transmission: the view from the slice. Pharmacology & Therapeutics. 2006;111:533–554. doi: 10.1016/j.pharmthera.2005.11.002. [DOI] [PubMed] [Google Scholar]
  182. Weiss F, Lorang MT, Bloom FE, Koob GF. Oral alcohol self-administration stimulates dopamine release in the rat nucleus accumbens: genetic and motivational determinants. Journal of Pharmacology and Experimental Therapeutics. 1993;267:250–258. [PubMed] [Google Scholar]
  183. Weiss F, Parsons LH, Schulteis G, Hyytia P, Lorang MT, Bloom FE, et al. Ethanol self-administration restores withdrawal-associated deficiencies in accumbal dopamine and 5-hydroxytryptamine release in dependent rats. Journal of Neuroscience. 1996;16:3474–3485. doi: 10.1523/JNEUROSCI.16-10-03474.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Weiss F, Porrino LJ. Behavioral neurobiology of alcohol addiction: recent advances and challenges. Journal of Neuroscience. 2002;22:3332–3337. doi: 10.1523/JNEUROSCI.22-09-03332.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Weitlauf C, Egli RE, Grueter BA, Winder DG. High-frequency stimulation induces ethanol-sensitive long-term potentiation at glutamatergic synapses in the dorsolateral bed nucleus of the stria terminalis. Journal of Neuroscience. 2004;24:5741–5747. doi: 10.1523/JNEUROSCI.1181-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Whittington MA, Little HJ. Nitrendipine, given during drinking, decreases the electrophysiological changes in the isolated hippocampal slice, seen during ethanol withdrawal. British Journal of Pharmacology. 1991;103:1677–1684. doi: 10.1111/j.1476-5381.1991.tb09846.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Wilson CJ, Kawaguchi Y. Th.e origins of two-state spontaneous membrane potential fluctuations of neostriatal spiny neurons. Journal of Neuroscience. 1996;16:2397–2410. doi: 10.1523/JNEUROSCI.16-07-02397.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Wise RA. Drug-activation of brain reward pathways. Drug and Alcohol Dependence. 1998;51:13–22. doi: 10.1016/s0376-8716(98)00063-5. [DOI] [PubMed] [Google Scholar]
  189. Witten IB, Lin SC, Brodsky M, Prakash R, Diester I, Anikeeva P, et al. Cholinergic interneurons control local circuit activity and cocaine conditioning. Science. 2010;330:1677–1681. doi: 10.1126/science.1193771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Wu JV, Kendig JJ. Differential sensitivities of TTX-resistant and TTX-sensitive sodium channels to anesthetic concentrations of ethanol in rat sensory neurons. Journal of Neuroscience Research. 1998;54:433–443. doi: 10.1002/(SICI)1097-4547(19981115)54:4<433::AID-JNR1>3.0.CO;2-A. [DOI] [PubMed] [Google Scholar]
  191. Xi ZX, Baker DA, Shen H, Carson DS, Kalivas PW. Group II metabotropic glutamate receptors modulate extracellular glutamate in the nucleus accumbens. Journal of Pharmacology and Experimental Therapeutics. 2002;300:162–171. doi: 10.1124/jpet.300.1.162. [DOI] [PubMed] [Google Scholar]
  192. Xia Y, Driscoll JR, Wilbrecht L, Margolis EB, Fields HL, Hjelmstad GO. Nucleus accumbens medium spiny neurons target non-dopaminergic neurons in the ventral tegmental area. Journal of Neuroscience. 2011;31:7811–7816. doi: 10.1523/JNEUROSCI.1504-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Xiao C, Shao XM, Olive MF, Griffin WC, 3rd, Li KY, Krnjevic K, et al. Ethanol facilitates glutamatergic transmission to dopamine neurons in the ventral tegmental area. Neuropsychopharmacology. 2009;34:307–318. doi: 10.1038/npp.2008.99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Xiao C, Ye JH. Ethanol dually modulates GABAergic synaptic transmission onto dopaminergic neurons in ventral tegmental area: role of muopioid receptors. Neuroscience. 2008;153:240–248. doi: 10.1016/j.neuroscience.2008.01.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Xie GQ, Wang SJ, Li J, Cui SZ, Zhou R, Chen L, et al. Ethanol attenuates the HFS-induced, ERK-mediated LTP in a dose-dependent manner in rat striatum. Alcoholism: Clinical & Experimental Research. 2009;33:121–128. doi: 10.1111/j.1530-0277.2008.00818.x. [DOI] [PubMed] [Google Scholar]
  196. Xu NL, Ye CQ, Poo MM, Zhang XH. Coincidence detection of synaptic inputs is facilitated at the distal dendrites after long-term potentiation induction. Journal of Neuroscience. 2006;26:3002–3009. doi: 10.1523/JNEUROSCI.5220-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Yim HJ, Gonzales RA. Ethanol-induced increases in dopamine extracellular concentration in rat nucleus accumbens are accounted for by increased release and not uptake inhibition. Alcohol. 2000;22:107–115. doi: 10.1016/s0741-8329(00)00121-x. [DOI] [PubMed] [Google Scholar]
  198. Yin HH, Park BS, Adermark L, Lovinger DM. Ethanol reverses the direction of long-term synaptic plasticity in the dorsomedial striatum. European Journal of Neuroscience. 2007;25:3226–3232. doi: 10.1111/j.1460-9568.2007.05606.x. [DOI] [PubMed] [Google Scholar]
  199. Zhang TA, Hendricson AW, Morrisett RA. Dual synaptic sites of D(1)-dopaminergic regulation of ethanol sensitivity of NMDA receptors in nucleus accumbens. Synapse. 2005;58:30–44. doi: 10.1002/syn.20181. [DOI] [PubMed] [Google Scholar]
  200. Zhou FM, Wilson CJ, Dani JA. Cholinergic interneuron characteristics and nicotinic properties in the striatum. Journal of Neurobiology. 2002;53:590–605. doi: 10.1002/neu.10150. [DOI] [PubMed] [Google Scholar]
  201. Zucca S, Valenzuela CF. Low concentrations of alcohol inhibit BDNF-dependent GABAergic plasticity via L-type Ca2+ channel inhibition in developing CA3 hippocampal pyramidal neurons. Journal of Neuroscience. 2010;30:6776–6781. doi: 10.1523/JNEUROSCI.5405-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES