ALCOHOL AND GENE EXPRESSION IN THE CENTRAL NERVOUS SYSTEM

TRAVIS J. WORST and KENT E. VRANA*

Center for the Neurobehavioral Study of Alcohol, Department of Physiology and Pharmacology, Wake Forest University School of Medicine, Winston-Salem, NC 27157, USA

* Author to whom correspondence should be addressed at: Department of Pharmacology, Penn State College of Medicine, 500 University Drive, Hershey, PA 17033, USA. Tel.: +1 717 531 8285; Fax: +1 717 531 0419; E-mail: kvrana{at}psu.edu

(Received 4 August 2004; first review notified 17 August 2004; in revised form 2 September 2004; accepted 24 October 2004)


    ABSTRACT
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
Aims: To describe recent research focusing on the analysis of gene and protein expression relevant to understanding ethanol consumption, dependence and effects, in order to identify common themes. Methods: A selective literature search was used to collate the relevant data. Results: Over 160 genes have been individually assessed before or after ethanol administration, as well as in genetically selected lines. Techniques for studying gene expression include northern blots, differential display, real time reverse transcriptase–polymerase chain reaction (RT–PCR) and in situ hybridization. More recently, high throughput functional genomic technology, such as DNA microarrays, has been used to examine gene expression. Recent gene expression analyses have dramatically increased the number of candidate genes (nine array papers have illuminated 600 novel gene transcripts that may contribute to alcohol abuse and alcoholism). Conclusions: Although functional genomic experiments (transcriptome analysis) have failed to identify a single alcoholism gene, they have illuminated important pathways and gene products that may contribute to the risk of alcohol abuse and alcoholism.


    INTRODUCTION
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
Although the causes of alcoholism are clearly multifactorial, human and animal studies have implicated genes in up to 60% of the overall risk (Buck et al., 1995Go; McGue et al., 1999Go; Dick and Foroud, 2003Go). Two central themes are shown in Figure 1. First, genetic predisposition suggests that a constellation of liability genes exist, which places an individual at risk for alcohol abuse or conversely protects an individual from risk. The possible genetic location of these genes has been addressed in several studies. One such example is the Collaboration on the Genetics of Alcoholism (COGA), a multi-site study involving clinically diagnosed human alcoholics and genetic associations, which indicates that there are four major chromosomal locations where such genes may reside (Reich et al., 1998Go). Traditionally, scientists have viewed this aspect of genetic risk as the potential for altered gene/protein function (cf. the decreased activity of the ALDH2*2 aldehyde dehydrogenase allele in Asian populations). However, a central premise of the present discussion is that such a genetic predisposition may also reside in altered levels of expression of a key gene. A modification of gene expression is the second component of risk that predisposes an individual towards alcoholism. In this case, chronic exposure to ethanol changes the patterns of gene expression in such a manner that it places the individual at risk for continued abusive drinking and ultimately, addiction. The normal homeostatic pattern of gene expression that characterizes the unperturbed nervous system assumes a new set point in the face of constant exposure to alcohol. This process of changing gene expression in response to chronic insult has been termed ‘allostasis’ (Sterling and Eyer, 1988Go) and may underlie important issues such as tolerance to the effects of alcohol. On balance then, the genetic components of alcohol abuse and alcoholism involve three different components (altered gene function, varying inherited patterns of gene expression and altered responses in gene expression to alcohol exposure). These genetic components of risk interact in a complicated manner with environmental factors. The latter two of these three aspects of genetic risk (both involving altered gene expression) are the subject of this review.



View larger version (13K):
[in this window]
[in a new window]
 
Fig. 1. Diagram of the genetic and environmental effects of ethanol. This figure illustrates the potential role of genetics and ethanol exposure in the development of a dependent state. The genome represents inherited gene expression in a ‘normal’ state. This state is altered upon exposure to alcohol, causing changes in expression that contribute to tolerance and/or dependence. After prolonged exposure, an allostatic change has occurred resulting in a dependant expression state. It is to be noted that genetic inheritance can predispose an individual to becoming dependent, just as environmental factors can contribute. The dotted lines represent issues that will not be directly addressed in this review, but those that are clearly of central importance to alcohol abuse and alcoholism.

 
This review focuses on the molecular biological aspects of gene regulation as reported in the literature. To this end, the text has been subdivided into neuropharmacological systems in an attempt to group together genes common to a specific system (e.g. pre- and post-synaptic receptors for a particular neurotransmitter). In an attempt to organize the material further, summaries that combine commonly studied themes for each system have also been provided. Although the summaries focus on repeatedly measured molecules, the reader should make conclusions based on all data presented.

With the wealth of knowledge available, bioinformatic interpretation is pivotal to a well-executed study. There are several software programs that can facilitate the completion of this task. A program maintained by the National Institutes of Health, DAVID (Database for Annotation, Visualization and Integrated Discovery; http://apps1.niaid.nih.gov/david/), offers several features such as ontological, pathway and functional classifications (Dennis et al., 2003Go). A program that deals specifically with pathways, GenMapp (Gene Microarray Pathway Profilier; http://www.genmapp.org/), has been created by the Gladstone Institutes of the University of California at San Francisco. Both programs can identify genes within a pathway or with common function that could illuminate new research directions or further clarify current results.

However, before delving into this literature, it is important that we recognize that at several points there are seemingly opposite findings or counter-intuitive results. This is particularly important in the small, but growing literature on DNA microarrays, where there are relatively few situations wherein the studies illuminate common genetic pathways. This problem arises from the fact that there are both direct pharmacological effects of ethanol, as well as long-term compensatory changes that occur in response to those pharmacological effects. These need not agree in an obvious way and so the route of administration, animal model and time course of a specific study will have a dramatic impact on the results.

Animals models and mode of administration
There are numerous animal models that have been employed to study the effects of alcohol. A unique research tool in the ethanol field, for instance, is the creation of genetically selected lines of varying ethanol phenotypes. Several such lines have been addressed within the following review, including the Alko Accepting and Alko Non-Accepting lines (AA and ANA, respectively) created in Finland (Eriksson, 1968Go) and the alcohol Preferring and Non-Preferring (P/NP) and High Alcohol Drinking and Low Alcohol Drinking (HAD/LAD) rats from the NIAAA-funded centre in Indianapolis (Lumeng et al., 1977Go; McBride and Li, 1998Go). A popular mouse model involves the comparison of C57Black/6 and DBA/2 mice, inbred strains that show differential ethanol preference and the recombinant inbred lines (BXD) created from an original F1 cross of the C57 and DBA animals (McClearn and Rodgers, 1959Go; Gora-Maslak et al., 1991Go). Similarly, two mouse lines have been employed to great benefit in the field. Withdrawal-seizure prone and resistant mice (WSP/WSR) differ in the seizure producing consequences of chronic ethanol administration (Crabbe and Phillips, 1993Go). In addition, the long sleep and short sleep (LS/SS) mice differ in their responses to the sedative effects of acute ethanol (McClearn and Kikihana, 1981Go).

A major area of conflict, when it comes to comparing disparate findings from individual reports, is the mode of ethanol administration. There are several distinct methods of administration that have differing effects on gene expression and it is important to realize that they have different pharmacological profiles and so might produce disparate results for a given genetic system. Such varying systems include intraperitoneal injection (IP), intragastric gavage (IG), oral self-administration (induced in a variety of distinct behavioural methodologies), inhalation chamber administration (with and without pharmacological manipulation of metabolism) and enforced administration as part of a liquid diet. The various approaches have differing levels of administration, varying levels of attendant administration stress, differing pharmacodynamic characteristics and unique behavioural contingencies. To illustrate this point, the differences in IP, IG and liquid diet have been studied and shown to cause increases in gene expression of NGFI-B using IP and IG but not liquid diet (Ogilvie et al., 1997aGo). Although the IP and IG doses were matched for dose, there was a significant difference in NGFI-B gene expression levels. Moreover, a liquid diet generating blood alcohol levels of 0.06–0.14% w/v failed to cause a change in NGFI-B levels as it had with IP and IG administration (Ogilvie et al., 1997aGo). It is therefore important to keep these issues in mind when comparing the results from different laboratories.

In general, this review of differential CNS gene expression in alcohol abuse and alcoholism is organized by neuropharmacological systems. Previously, studies have focused on those systems thought to be directly impacted by alcohol—GABA, glutamate, opiates and dopamine. Therefore, there is a wealth of information evolving for these systems. Growing evidence also implicates signal transduction/transcription factor systems. Finally, we examine the wealth of information that is beginning to emerge from the use of multiplex DNA hybridization arrays (DNA macroarrays, microarrays and GeneChips). As an added bibliographic tool, all of the genes discussed in the review are provided in a public website with associated PubMed reference links and NCBI-supported genetic LocusLink (www.arraydata.org/AlcoholReview).


    GABA SYSTEM
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
There has long been an association between gamma-aminobutyric acid (GABA) and ethanol (Grant and Lovinger, 1995Go; Davis and Wu, 2001Go). The GABA system receives considerable focus because of similarities in the pharmacological effects of ethanol and GABA agonists, characterization of an ethanol-binding site on the GABA receptor and the fact that several different GABA receptor subunits show robust acute and chronic responses to ethanol exposure. Receptor subunit expression makes up a large amount of ethanol-associated GABA research. A current summary of gene expression changes in GABAergic receptors is presented in Table 1. The GABAA receptor is comprised of a combination of five of the following subunits: {alpha}, ß, {delta}, {gamma} and {varepsilon} (for a review, see Costa, 1998Go). The {alpha}1 GABAA subunit has been the subject of several studies, such as two initial reports showing a decrease in expression in the cortex (Buck et al., 1991Go; Montpied et al., 1991Go). This finding was confirmed in several other laboratories and models (Mhatre and Ticku, 1992Go; Devaud et al., 1995Go, 1997Go; Matthews et al., 1998Go; Eravci et al., 2000Go). One study has shown an increase in {alpha}1 expression, with an average blood ethanol concentration (BEC) of 250 mg/dl in mice treated for 7 days and allowed to withdraw for 8 h (Hirouchi et al., 1993Go). Interestingly, the lone post-mortem study of GABAA {alpha}1 in human alcoholics showed no change in the gene expression levels (Mitsuyama et al., 1998Go).


View this table:
[in this window]
[in a new window]
 
Table 1. Summary of major neurotransmission-associated genes

 
The {alpha}2 and {alpha}3 GABA receptor subunits have been analysed in a complex model of ethanol consumption designed to examine long-term changes in gene expression responses to varying contexts of ethanol administration. These studies used four groups of animals: controlled consumers, behaviourally dependent consumers, forced consumers and control animals (Eravci et al., 1997Go). Controlled consumer rats were given water as the sole fluid for 12 months. In the next month there was a choice between water, 5% and 20% ethanol, whereas the following month had 5% and 20% ethanol spiked with 0.1 g/l quinine or water as three choices. These control animals exhibited modest drinking of alcohol (1.1 g/kg/day) that was severely decreased by the aversive quinine (0.4 g/kg/day). This assessment period was followed, for all groups, by a month's abstinence prior to sacrifice. The dependent consumers differed from the above protocol by having 9 months experience of free access to water, 5% and 20% ethanol, followed by 3 months of water as the sole fluid (enforced abstinence), with the final 3 months identical to the controlled consumers (including a month's abstinence). These animals showed high levels of drinking (3.1 g/kg/day) and were behaviourally dependent in that they consumed 2.1 g/kg/day in the face of quinine addition. The forced consumer group is identical to the behaviourally dependent group, except for the first 9 months wherein 5% ethanol was the sole fluid and engendered modest drinking following abstinence (1.9 g/kg/day). The concept of this administration protocol is based on choice of self-administration. The important measure is the amount of free-choice ethanol consumed for 1 month after water, self-administration of ethanol or forced ethanol consumption with a 3 month withdrawal period in between. Indeed, another salient feature of this protocol is that all gene expression was assessed following 1 month of ethanol and so reflects robust and enduring effects. The {alpha}2 and {alpha}3 subunits demonstrated decreases in gene expression in controlled consumers and behaviourally dependent rats for both subunits, whereas the forced consumers showed a decrease only in the {alpha}3 subunit (Eravci et al., 2000Go). This differs from a study from the Biggio laboratory showing an increase in {alpha}2 expression in a primary culture of cerebellar neurons treated for five days with 100 mM ethanol (Follesa et al., 2004Go), arguing that the context of the ethanol administration plays a role in the molecular consequences. The Eravci model will be revisited for additional gene targets in subsequent discussions.

Buck et al. (1991)Go found changes in GABAA receptor mRNA expression in WSP and WSR mice prior to and after ethanol administered in a liquid diet. In ethanol-naïve comparisons, the WSP mice demonstrate a decrease in {alpha}3 expression versus WSR strains. In contrast, after ethanol treatment, both lines show upregulation of the {alpha}3 subunit compared with ethanol-naïve animals. These investigators conclude that the {alpha}3 subunit may play a small role compared with the increased gene expression of the {alpha}1 subunit in contributing to the genetic difference between the two strains.

An in situ study of the {alpha}4 subunit reported changes in mRNA expression in several different brain regions, such as an increase in the thalamus, layers 2 and 3 of the cortex and the CA1, CA3 regions and dentate gyrus of hippocampus following 5 g/kg/day of IG ethanol (Mahmoudi et al., 1997Go). Another {alpha}4 subunit study exposed rats to 5% ethanol in a liquid diet for 1 week, then 7.5% ethanol for 1 week followed by 1 day of withdrawal (Davis-Cox et al., 1996Go). The animals demonstrated increases in {alpha}4 gene expression in the cortex as assessed by reverse transcriptase–polymerase chain reaction (RT–PCR). Protein levels were increased in the cortex and hippocampus of rats after a BEC of 250 mg/dl over a range of 14–40 days and withdrawal periods of 0–24 h (Devaud et al., 1997Go; Matthews et al., 1998Go). In contrast to these findings, a post-mortem human study found no change in {alpha};4 gene expression using RT–PCR (Mitsuyama et al., 1998Go). All studies, with the exception of this human post-mortem investigation, reported an increase in expression for the {alpha}4 subunit gene. This type of consistency in results from different laboratories is relatively uncommon in the gene expression field and indicates a potentially important role of the {alpha}4 subunit. At the same time, however, the Mitsuyama et al. (1998)Go study needs confirmation. If replicated, these human results force a re-examination of the relationships with rodent models.

The {alpha}5 and {alpha}6 subunits were both analysed as components of larger studies and showed differential expression in various regions of the brain. Eravci et al. (2000)Go show downregulation of the {alpha}5 subunit in controlled consumers and behaviourally dependent rats (even following a month of enforced abstinence). The {alpha}6 subunit was decreased in whole brain of ethanol-naïve WSP compared with WSR mice, but WSR mice showed a decrease after ethanol treatment (Buck et al., 1991Go). Finally, one study demonstrated a decrease in {alpha}5 subunit expression in the cortex after IG administration of 20% ethanol thrice daily for 6 days after both 1 and 24 h withdrawal, but an increase in {alpha}6 in the cerebellum was found on following the same protocol (Mhatre and Ticku, 1992Go).

The ß subunits have had less attention but show gene expression responses to ethanol. The ß2 subunit mRNA expression was decreased in ethanol-naïve C57Black mice compared with DBA2 and treated C57Black mice compared with naïve C57Black mice in whole brains at low ethanol concentrations, whereas higher concentrations of ethanol showed an increase. The DBA2 mice showed an increase at intermediate and higher doses of ethanol compared with ethanol-naïve DBA2 mice (Reilly and Buck, 2000Go). One study noted decreases in the ß1 subunit in a behaviourally dependent cohort and decreases in the ß2 subunit in controlled consumers as well as behaviourally dependent animals (Eravci et al., 2000Go), whereas others demonstrated increases in ß2 and ß3 subunit protein expression within the cortex after 14 days of a 6–7.5% liquid diet (Devaud et al., 1997Go, 1998Go). Another study, involving the ß3 subunit in human post-mortem samples, showed an increase in gene expression within the frontal cortex (Mitsuyama et al., 1998Go).

The {gamma} subunits have received more attention since the discovery that the {gamma}2 short and long variants may have differential expression (Devaud et al., 1995Go). Interestingly, {gamma}2s and {gamma}1 mRNA expression were found increased in cortex after 2 weeks of an ethanol-containing diet (Devaud et al., 1995Go). Later, Devaud et al. (1997)Go analysed {gamma}1 protein expression in the cortex, discovering an upregulation 0 and 8 h after removal from an ethanol liquid diet. Several studies have shown a decrease in gene expression in both the {gamma}2l and {gamma}2s subunits in the parieto-occipital cortex (POC) and cortex (Eravci et al., 2000Go; Mhatre and Ticku, 1992Go, respectively). Expression of {gamma}2l mRNA has also been reported as decreased in the dendritic and pyramidal layers of the CA1 region of the hippocampus (Petrie et al., 2001Go).

The inhibitory GABA receptor is noteworthy for its complex modulation by several molecules, including the diazepam binding inhibitor (DBI). DBI acts as an inverse agonist that binds to the benzodiazepine site on GABA receptors (Ohkuma et al., 2001Go). One study focused on DBI in cultured cerebral cortical neurons of mice (Katsura et al., 1995aGo). This study had two parts: the first experiment held the concentration of ethanol at 50 mM for 1–5 days, the second used concentrations of ethanol from 1 to 100 mM for 3 days. The end result was an increase in expression at all points for the first experiment and all but the 1 mM concentration for the second. Coincidently, this laboratory published whole animal studies in mice with an average BEC of 250 mg/dl over 8 days and showed that DBI mRNA was increased after 0 and 8 h of withdrawal (Katsura et al., 1995bGo). A subsequent experiment by this group involved a project having a well-balanced mix of mRNA and protein level measures, conducted in mice subjected to inhalation chamber administration for 8 days, producing a BEC of ~250 mg/dl (Katsura et al., 1998Go). They found increases in gene expression after no withdrawal, 8 h, 1, 2, 3, 4, 5 and 7 days following the last treatment. Protein levels displayed a similar pattern beginning at the 8 h point—this latter finding would be consistent with a functional lag between changes in mRNA and protein.

In summary, of the several subunits studied to date, the {alpha}4 and {gamma}2l subunits appear to be the most consistently regulated, indicating that they are of fundamental importance. These two subunits alone are the subject of eight individual studies covering 11 separate regions of the brain. The {alpha}4 subunit has been loosely associated with tolerance (Mahmoudi et al., 1997Go), such that increasing this subunit may represent an attempt to normalize the genomic effects of alcohol. The {gamma}2l subunit has been associated with protein kinase C (PKC) and phosphorylation, indicating that ethanol may alter phosphorylation states and second messenger activity, possibly leading to increased receptor function. On the other hand, the GABAergic system is also very plastic, showing differential expression responses in numerous subunits, thereby making these subunits prime targets for development of pharmacotherapies. In addition, the inhibitory nature of GABA is in contrast to the excitatory nature of the glutamate system (discussed below) and provides an ideal environment for investigating a cause and effect relationship between these two systems.


    GLUTAMATE SYSTEM
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
Most researchers have approached the study of glutamate from the angle of receptor expression as summarized in Table 1. Glutamate receptors are divided into two classes—ionotropic and metabotropic (Meldrum, 2000Go). The ionotropic receptors are classified as alpha-amino-3-hydroxy-5-methyl-4-isoxazole-propionic acid (AMPA), kainite or N-methyl-D-aspartate (NMDA): they are channels allowing the flow of ions through the cell membrane. Eravci et al. (2000)Go describe a decrease in AMPA receptor 2 and kainate receptor 2 in the POC of rats (Eravci et al., 2000Go). Gene expression levels of the AMPA receptor were decreased in the POC of controlled consumers, whereas controlled consumers as well as behaviourally dependent rats showed decreases in kainate receptor gene expression. Once again, the controlled consumers are animals that had access only to water for 12 months prior to free-choice access to ethanol or water and then a 1 month period of abstinence. On the other hand, Rimondini et al. (2002)Go found an increase in AMPA receptor 2 after 3 weeks withdrawal from 7 weeks of 150–250 mg/dl of ethanol in an inhalation chamber. This change was noted in the cingulate cortex of the rats tested. In the same model, AMPA receptor interacting protein was increased within the amygdala.

The NMDA-sensitive receptors are, by far, the most frequently associated with ethanol effects. In the previously noted study with AMPA and kainate receptors in the POC, Eravci et al. (2000)Go reported decreases in the mRNA levels for NMDA1 and 2a subunits in the controlled consumers. Additionally, the 2a subunit was also decreased in the behaviourally dependent model. These animals had free access to ethanol or water for 9 months prior to 3 months abstinence, a 1 month testing period and 1 further month of abstinence. This depression agrees with Feuerstein and collaborators, who have shown a decrease in the 2b subunit in the CA region of the hippocampus as well as caudate putamen (Darstein et al., 2000Go). Two groups have shown increases in 2b subunit expression: one demonstrated an increase in 2b mRNA in cultured mouse cortical neurons after 5 days of exposure to 50 mM ethanol (Hu et al., 1996Go); the second showed that an inhalation chamber paradigm, producing a BEC of 200–400 mg/dl, also increased 2b mRNA within the cortex (Hardy et al., 1999Go). Moving to the 2d subunit, Naassila and Daoust (2002)Go reported increases in gene and protein expression in the hippocampus in pups of ethanol-drinking dams. However, in a study of a selected line, ethanol-naïve Alcohol Tolerant (AT) rats showed no difference from Alcohol Non Tolerant (ANT) rats in 2a, 2b, 2c and 2d NMDA subunit gene expression (Toropainen et al., 1997Go). This change was also noted in protein expression.

The NMDA1 subunit has also been the focus of significant attention. Characterization of ethanol-naïve AA versus ANA rats demonstrated an increase in subunit variant 1-4 within the hippocampus of ANA rats (Winkler et al., 1999Go). This study also reported an increase in protein expression in the hippocampus from the alternative splice variant C-terminus subunit 1-3/1-4 after ethanol treatment in AA rats. A fetal alcohol syndrome model demonstrated increases in subunit 10xx in the cerebellum 1, 7 and 14 days post-parturition and 7 and 14 days in the hippocampus for both 10xx and 11xx splice variants (Naassila and Daoust, 2002Go). Hardy et al. (1999)Go reported an increased expression of the 10xx over the 11xx within the cortex at 0, 6, 12, 24 and 48 h after inhalation chamber administration of ethanol.

Receptors are not the only molecules that affect glutamatergic function—enzymes and transporters can also have an impact on the activity of this neurotransmitter system. Within the POC, mitochondrial and cytosolic aspartate aminotransferase gene expression was decreased in both controlled consumers and behaviourally dependent models (Eravci et al., 1999Go). This gene encodes a glutamate converting enzyme, changing glutamate into {alpha}-ketoglutarate and vice versa, although the thermodynamic balance is higher for degrading glutamate (McKenna et al., 2000Go). Two other glutamate degrading enzymes include glutaminase and glutamine synthetase (Svenneby and Torgner, 1987Go). Glutaminase was decreased in the behaviourally dependent rats in the POC, along with glutamine synthetase in both controlled consumers and behaviourally dependent rats. A hybridization array experiment also revealed an increased expression of the glutamate/aspartate transporter in the cingulate cortex (Rimondini et al., 2002Go). Eravci and colleagues tested other glutamate-related enzymes as well, and showed a decrease in glutamate dehydrogenase in the POC in both ethanol self-administering models as well as a decrease in a GABA producing enzyme, glutamic acid decarboxylase 65 (GAD65), in the behaviourally dependent model. On the other hand, GAD65 expression is increased in the controlled consumers and behaviourally dependent rats within the limbic forebrain (Eravci et al., 1999Go). Two other unique NMDA receptor-like molecules, glutamate- and glycine-binding subunits, originally discovered in synaptic membrane preparations, were demonstrated to be increased in cortical cultures after 3 days of treatment with 100 mM ethanol (Bao et al., 2001Go).

Similar to the GABAergic system, the glutamate system demonstrates more robust receptor changes rather than ligand or metabolic effects. Examples include increased gene expression in nearly all NMDA receptor subunits, such as NR1-3, NR1-4, NR2A, NR2B and NR2D but with few changes in glutamate degrading enzymes, such as decreases in glutaminase, glutamine synthase and GAD65. This may reflect the ubiquitous nature of glutamate, requiring that regulatory responses reside within the receptors. Moreover, the common theme of receptor regulation within GABAergic and glutamatergic systems, along with the unique relationship of GABA activity robustly opposing glutamate function in the presence of ethanol, make these two systems particularly interesting.


    OPIATE SYSTEM
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
The opiate system has long been implicated in alcoholism, owing in part to the similar anaesthetic/depressant properties of the two classes (Herz, 1997Go). Hence, several opiate receptors and precursors of endogenous opiates have been extensively studied at the level of mRNA expression (Table 1). The mRNAs for the µ, {kappa}, and {delta} receptors, as well as the precursors to Pro-dynorphin and Pro-enkephalin [prepro-dynorphin (PPD) and prepro-enkephalin (PPE), respectively] and ß-endorphin have all been investigated. ß-endorphin, encoded by pro-opiomelanocortin (POMC), has an equal affinity for µ and {delta} receptors. Dynorphin is specific for the {kappa} receptors, while met- and leu-enkephalin have a higher affinity for {delta} receptors, but also bind to µ receptors. Not only do the receptors differ in ligand specificity but also in localization (Miotto et al., 1996Go), as well as locomotor and behaviour effects (Kieffer and Gaveriaux-Ruff, 2002Go).

The µ and {delta} receptors were both evaluated in several regions of the brain in a single paper by Winkler et al. (1998b)Go beginning with comparisons in ethanol-naïve C57Black versus DBA2 mice (Winkler et al., 1998bGo). These two mouse lines exhibit markedly different sensitivities and self-administration behaviours for ethanol. These investigators reported no changes in gene expression level between the two strains. However, using the C57Black, they showed a decrease in the µ receptor in the hypothalamus after ethanol treatment. The mice self-administered 10% ethanol for 28 days consuming average values of ethanol of 8.8 and 3 g/kg/day for the C57Black and DBA2 mice, respectively. RT–PCR was performed after withdrawal periods of 0, 3 and 21 days to evaluate mRNA levels for the opiate receptors. The DBA2 mice also showed a decrease in the hypothalamus after no withdrawal, but not after 3 or 21 days post ethanol. Winkler addressed the {kappa} receptor in a separate study and showed an increase receptor mRNA expression in the DBA2 mice in the regions of the hypothalamus and striatum (Winkler et al., 1998aGo). The {kappa} and µ receptors were also increased in the ANA compared with AA and Wistar rats as assessed by receptor binding studies, whereas RT–PCR demonstrated increases in pro-enkephalin, pro-dynorphin, and POMC (Marinelli et al., 2000Go).

In another series of experiments, Jamensky and Gianoulakis (1999)Go reported an increase in naïve C57Black versus DBA2 opiate system mRNA expression levels using in situ hybridization. The increases were noted for POMC in the arcuate hypothalamic nucleus and pro-enkephalin in the caudate putamen and nucleus accumbens. POMC was also increased after an IP administration of ethanol (Krishnan-Sarin et al., 1998Go; Kinoshita et al., 2000Go) and administration through a liquid diet (Angelogianni et al., 1993Go; Chen et al., 2004Go). Contrary to the Krishnan-Sarin studies, several groups have demonstrated decreases in POMC following an inhalation chamber paradigm, liquid diet and intragastric administration (Scanlon et al., 1992Go; Aird et al., 1997Go; Zhou et al., 2000Go). These studies demonstrate that POMC regulation is dependent on the type of administration and the region of the brain, but is likely to play an important physiological role in response to ethanol.

In a series of experiments led by de Gortari et al. (2000)Go, rats were intragastrically administered once with 2.5 g/kg of ethanol and allowed to withdraw for 1, 4 or 24 h after the single injection. Levels of pro-enkephalin were found decreased in hippocampal regions at the early time point and were reduced at longer withdrawal periods in the frontal cortex and nucleus accumbens. However, enkephalin mRNA was increased at later time points in hippocampal regions, mammillary bodies of the hypothalamus and nucleus accumbens. Another pro-enkephalin study showed no change in expression using in situ hybridization in the rostral striatum after chronic ethanol consumption (Tajuddin and Druse, 1998Go). Pro-dynorphin, on the other hand, was increased in whole brains of withdrawal-seizure prone mice 6 h after ethanol inhalation (Beadles-Bohling et al., 2000Go). The mice had an average BEC of 150 mg/dl for the 3 day treatment period. Plotkin et al. (1997)Go conducted a whole-brain study that examined met-enkephalin and PPE in mice treated with a 5% ethanol liquid diet for 10 days. Although the PPE mRNA levels did not change, the met-enkephalin protein levels measured by radioimmunoassay, increased immediately after treatment, at 10 h and at 8 days post treatment. Concussion and ethanol consumption was the model used to query PPE mRNA levels after a 2.4 g/kg IP dose of ethanol (Sall et al., 1996Go). These investigators set out to test the hypothesis that alcohol consumption alters PPE levels, leading to increased neuronal injury and death. Prior to concussion, PPE levels in the frontal cortex and olfactory bulb were found to be decreased 24 h after the neuronal injury. PPE was decreased in the concussion model with ethanol injection 10 min prior to injury in piriform/amygdala cortex and also increased in the same model in the olfactory bulb. This change within the olfactory bulb is opposite to the regulation seen in the ethanol-only treated rats, indicating the influence of trauma in that region. A comparison of naïve ethanol-preferring Fawn-Hooded and non-preferring Kyoto–Wistar rats revealed an increase in PPD in the hippocampus and a decrease in PPE in the nucleus accumbens and striatum, as seen using in situ hybridization (Cowen et al., 1998Go).

An additional PPE study analysed P and NP rats that are ethanol-naïve or treated with 2.5 g/kg of ethanol (IG) and allowed to withdraw for 1, 4 or 8 h (Li et al., 1998Go). PPE mRNA was found to be increased in ethanol-naive P over NP rats in the anterior hypothalamus, as well as the posterior and medial striatum, 4 h after saline treatment. After 8 h, the increase was seen in the medial shell of the nucleus accumbens. Four hours after ethanol treatment, PPE was increased again in the posterior striatum and the amygdala in the P versus the NP lines. Within-line comparisons were also performed. The P rats showed increased PPE levels 1 h after ethanol exposure in three regions of the nucleus accumbens: the whole accumbens, the lateral core and the medial shell. Eight hours after exposure, the naïve P rats had higher PPE levels in the medial shell and whole accumbens. In a similar manner, 8 h after treatment, the naïve NP rats had higher gene expression than exposed NP rats in the lateral core, medial shell and whole accumbens.

Unlike the inhibitory (GABA) and excitatory (glutamate) neurotransmitter systems discussed above, the opiates tend to have more emphasis on precursor molecules rather than receptor changes. This is reflected by an increase observed in 62% of the ligands tested in several regions of the brain. Presumably, the change in ligand amount would directly reflect action at receptors, leading to increased or decreased opioid function.


    ALTERNATIVE NEUROTRANSMITTER, HORMONE AND TRANSCRIPTION FACTOR SYSTEMS
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
Dopamine and serotonin have been implicated in ethanol abuse, as well as many other forms of abuse, and have been subject to extensive study. Specifically, dopamine levels are increased by nearly all drugs of abuse, whereas serotonin is closely associated with impulsive behaviour (for a review of both, see Lingford-Hughes et al., 2003Go).

The serotonin 1a and 1b receptors were investigated at the level of mRNA and protein (Nevo et al., 1995Go). The 1a receptor mRNA demonstrated no change 18 h after withdrawal following 2 weeks of a 10% ethanol liquid diet and an additional 5 g/kg of IG dose of ethanol. The 1a receptor protein, however, increased expression (based on autoradiography studies) in the dorsal raphe and decreased expression in the dentate gyrus, entorhinal cortex and layer IV of the frontal cortex. The same investigators went on to show an increase in serotonin 1b receptor mRNA expression within the striatum and protein expression within the globus pallidus.

The rate-limiting step of dopamine biosynthesis is tyrosine hydroxylase (TH). In fact, as the key regulator for all catecholamine synthesis, this enzyme (and its gene) has been shown to be subject to nearly every form of regulation (Kumer and Vrana, 1996Go). Szot et al. (1999)Go have demonstrated a decrease in gene expression for TH in the ventral tegmental area of offspring of dams treated on day 21 of pregnancy through birth with 12 g/kg of ethanol in a liquid diet. This is intriguing considering that theories of dopamine-related reward revolve around the change in dopamine access or production in the synapse as seen in cocaine studies (Drago et al., 1998Go). The dopamine D3 receptor exhibits decreased expression in the limbic forebrain in controlled consumers and behaviourally dependent rats and in the hippocampus of controlled consumers (Eravci et al., 1997Go). Dopamine transporter gene expression in the ventral tegmental area and substantia nigra pars compacta is also decreased in the offspring of rats treated with 12 g/kg of ethanol from day 21 through the end of pregnancy (Szot et al., 1999Go).

As summarized above, the monoaminergic transmitter systems demonstrate mixed responses, although both dopamine and serotonin systems show a trend for downregulation of gene expression. The few studies on these systems also make it difficult to draw conclusions about an overall effect of ethanol. However, there is consistent downregulation of the dopamine D3 autoreceptor and dopamine transporter, indicating a possible method for increasing synaptic dopamine.

In addition to these alternative neurotransmitter systems, a variety of hormone and transcription factors have been implicated in alcohol abuse and alcoholism. In the interests of space, these genes and gene products are summarized in Tables 1 and 2.


View this table:
[in this window]
[in a new window]
 
Table 2. Summary of miscellaneous genes

 

    ARRAY MANUSCRIPTS
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
The latest development in functional genomics is high throughput screening (DNA microarray, or Gene Chip). This technology has only recently been applied to the field of ethanol research, in that there are nine references until mid 2004 (Lewohl et al., 2000Go; Thibault et al., 2000Go; Xu et al., 2001Go; Daniels and Buck, 2002Go; Mayfield et al., 2002Go; Rimondini et al., 2002Go; Saito et al., 2002Go; Sokolov et al., 2003Go; Treadwell and Singh, 2004Go). One is a re-examination of samples from a single cohort of human alcoholics. There is remarkably little commonality illuminated by these different studies (Table 3).


View this table:
[in this window]
[in a new window]
 
Table 3. Summary of DAVID Gene Ontology search of array data

 
Perhaps the most striking of the array manuscripts are the human post-mortem studies by Lewohl et al. (2000)Go, Mayfield et al. (2002)Go and Sokolov et al. (2003)Go. All studies illuminate biological themes, such as the ethanol-induced regulation of several myelin-associated genes noted by Lewohl et al. (2000)Go. Although the Lewohl et al. (2000)Go and Mayfield et al. (2000) reports originate from the same laboratory (but use differing samples from the same source of human post-mortem tissue), only two genes complement each other out of the 169 reported by the former study and the 70 by the latter. These two studies did differ, however, in their patient inclusion criteria. In addition, Sokolov et al. (2003)Go report changes in expression of 173 genes within the temporal cortex of human post-mortem alcoholics. Taken together, only six genes are commonly regulated in two of the three studies and no genes are present in all three studies combined. Unfortunately, these six common genes do not naturally lend themselves to a common biological narrative. Lewohl et al. (2000)Go and Mayfield et al. (2002)Go conclude that changes in myelin expression may explain the loss of white matter and demyelination that occurs in alcoholics, whereas Sokolov et al. (2003)Go focus more on calcium regulation and energy metabolism.

Thibault et al. (2000)Go have also provided evidence, with cAMP and ethanol demonstrating a similar regulation on a subset of neuroblastoma genes responsive to ethanol. Selected targets from this work were confirmed using other methods, such as RT–PCR, which is proving to be the normal practice in the array field. A single myelin-associated gene, myelin basic protein, appeared in other array work. This is not surprising considering the use of cultured human neuroblastoma cells in this manuscript versus the human post-mortem samples used by Lewohl et al. (2000)Go and Mayfield et al. (2002)Go.

Saito et al. (2002)Go moved to the rat model and identified another 165 differentially-expressed genes. This laboratory identified two general physiological systems that are affected by ethanol—oxidative stress and membrane trafficking. This study was performed on hippocampal regions and in an entirely different species, so the lack of matching expression patterns is not necessarily a point of concern.

Another array format manuscript, performed by Rimondini et al. (2002)Go, reported changes in rats after three weeks of exposure to ethanol. They focused on the cingulate cortex and amygdala, noting changes of gene expression in three main areas—neurotransmission, synaptic plasticity and signal transduction. The difference in ethanol treatment and administration may be responsible for the lack of replicated genes between this and the study of Saito et al. (2002)Go discussed previously.

The message regarding array work is two-fold: (i) several models can independently demonstrate a large number of changes in gene expression; and (ii) different models are important for ethanol research but may identify different results. The different models will each give details on specific aspects of the disease, and large-scale gene studies reflect changes in those aspects. Therefore, inconsistent data, although troubling, increases the base of knowledge that others may build upon.


    CONCLUSIONS
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
The studies presented above cover less than two decades of investigation, but the amount of data is staggering. There is obvious involvement of the GABAergic and glutamatergic neurotransmitter systems, as well as serotonin and dopamine. Not only are receptors and receptor subunits affected, but neurotransmission machinery and architecture could also impact neurotransmitter release or lack thereof. The advent of array platforms has already added to the knowledge base of the field, with new gene targets being reported each month. With the completion of the genome sequencing projects, researchers may be able to correlate changes in mRNA or protein expression to splice variants or mutations in DNA sequences. A coherent story has yet to take shape. Evolving techniques and models, elucidation of roles for current genes and discovery of new genes combined with a growing bioinformatic discipline, provide hope that a clearer understanding of the functional genomic effects of ethanol is within reach.


    ACKNOWLEDGEMENTS
 
This research has been supported by National Institute of Health grants P50-AA1197 (K.E.V.) and T32-AA07565 (T.J.W.).


    REFERENCES
 TOP
 ABSTRACT
 INTRODUCTION
 GABA SYSTEM
 GLUTAMATE SYSTEM
 OPIATE SYSTEM
 ALTERNATIVE NEUROTRANSMITTER,...
 ARRAY MANUSCRIPTS
 CONCLUSIONS
 REFERENCES
 
Acquaah-Mensah, G. K., Leslie, S. W. and Kehrer, J. P. (2001) Acute exposure of cerebellar granule neurons to ethanol suppresses stress-activated protein kinase-1 and concomitantly induces AP-1. Toxicology, and Applied Pharmacology 175, 10–18.[CrossRef][ISI][Medline]

Aird, F., Halasz, I. and Redei, E. (1997) Ontogeny of hypothalamic corticotropin-releasing factor and anterior pituitary pro-opiomelanocortin expression in male and female offspring of alcohol-exposed and adrenalectomized dams. Alcoholism: Clinical and Experimental Research 21, 1560–1566.[ISI][Medline]

Alfos, S., Higueret, P., Pallet, V., Higueret, D., Garcin, H. and Jaffard, R. (1996) Chronic ethanol consumption increases the amount of mRNA for retinoic acid and triiodothyronine receptors in mouse brain. Neuroscience Letters 206, 73–76.[CrossRef][ISI][Medline]

Alfos, S., Boucheron, C., Pallet, V., Higueret, D., Enderlin, V., Beracochea, D., Jaffard, R. and Higueret, P. (2001) A retinoic acid receptor antagonist suppresses brain retinoic acid receptor overexpression and reverses a working memory deficit induced by chronic ethanol consumption in mice. Alcoholism: Clinical and Experimental Research 25, 1506–1514.[CrossRef][ISI][Medline]

Angelogianni, P. and Gianoulakis, C. (1993) Chronic ethanol increases proopiomelanocortin gene expression in the rat hypothalamus. Neuroendocrinology 57, 106–114.[ISI][Medline]

Anokhina, I. P., Ovchinnikova, L. N., Shamakina, I. Y., Khristolyubova, N. A., Krylova, O. Y. and Gorkin, V. Z. (1990) Some neurobiological mechanisms of the effect of ethanol on offspring of chronically alcohol treated rats. Annals of Medicine 22, 353–356.[ISI][Medline]

Bachtell, R. K., Wang, Y. M., Freeman, P., Risinger, F. O. and Ryabinin, A. E. (1999) Alcohol drinking produces brain region-selective changes in expression of inducible transcription factors. Brain Research 847, 157–165.[CrossRef][ISI][Medline]

Baek, J. K., Heaton, M. B. and Walker, D. W. (1994) Chronic alcohol ingestion: nerve growth factor gene expression and neurotrophic activity in rat hippocampus. Alcoholism: Clinical and Experimental Research 18, 1368–1376.[ISI][Medline]

Baek, J. K., Heaton, M. B. and Walker, D. W. (1996) Up-regulation of high-affinity neurotrophin receptor, trk B-like protein on western blots of rat cortex after chronic ethanol treatment. Brain Research. Molecular Brain Research 40, 161–164.[ISI][Medline]

Bao, X., Hui, D., Naassila, M. and Michaelis, E. K. (2001) Chronic ethanol exposure increases gene transcription of subunits of an N-methyl-D-aspartate receptor-like complex in cortical neurons in culture. Neuroscience Letters 315, 5–8.[CrossRef][ISI][Medline]

Beadles-Bohling, A. S., Crabbe, J. C. and Wiren, K. M. (2000) Elevated prodynorphin expression associated with ethanol withdrawal convulsions. Neurochemistry International 37, 463–472.[CrossRef][ISI][Medline]

Breese, C. R., D'Costa, A., Ingram, R. L., Lenham, J. and Sonntag, W. E. (1993) Long-term suppression of insulin-like growth factor-1 in rats after in utero ethanol exposure: relationship to somatic growth. Journal of Pharmacology and Experimental Therapeutics 264, 448–456.[Abstract]

Breese, C. R., D'Costa, A. and Sonntag, W. E. (1994) Effect of in utero ethanol exposure on the postnatal ontogeny of insulin-like growth factor-1, and type-1 and type-2 insulin-like growth factor receptors in the rat brain. Neuroscience 63, 579–589.[CrossRef][ISI][Medline]

Buck, K. J. (1995) Strategies for mapping and identifying quantitative trait loci specifying behavioral responses to alcohol. Alcoholism: Clinical and Experimental Research 19, 795–801.[ISI][Medline]

Buck, K. J., Hahner, L., Sikela, J. and Harris, R. A. (1991) Chronic ethanol treatment alters brain levels of gamma-aminobutyric acidA receptor subunit mRNAs: relationship to genetic differences in ethanol withdrawal seizure severity. Journal of Neurochemistry 57, 1452–1455.[ISI][Medline]

Caberlotto, L., Thorsell, A., Rimondini, R., Sommer, W., Hyytia, P. and Heilig, M. (2001) Differential expression of NPY and its receptors in alcohol-preferring AA and alcohol-avoiding ANA rats. Alcoholism: Clinical and Experimental Research 25, 1564–1569.[CrossRef][ISI][Medline]

Chen, W., Hardy, P. and Wilce, P. A. (1997) Differential expression of mitochondrial NADH dehydrogenase in ethanol-treated rat brain: revealed by differential display. Alcoholism: Clinical and Experimental Research 21, 1053–1056.[ISI][Medline]

Chen, C. P., Kuhn, P., Advis, J. P. and Sarkar, D. K. (2004) Chronic ethanol consumption impairs the circadian rhythm of pro-opiomelanocortin and period genes mRNA expression in the hypothalamus of the male rat. Journal of Neurochemistry 88, 1547–1554.[CrossRef][ISI][Medline]

Costa, E. (1998) From GABAA receptor diversity emerges a unified vision of GABAergic inhibition. Annual Review of Pharmacology and Toxicology 38, 321–350.[CrossRef][ISI][Medline]

Cowen, M. S., Rezvani, A., Jarrott, B. and Lawrence, A. J. (1998) Distribution of opioid peptide gene expression in the limbic system of Fawn-Hooded (alcohol-preferring) and Wistar–Kyoto (alcohol-non-preferring) rats. Brain Research 796, 323–326.[CrossRef][ISI][Medline]

Crabbe, J. C. and Phillips, T. J. (1993) Selective breeding for alcohol withdrawal severity. Behavior Genetics 23, 171–177.[ISI][Medline]

Daniels, G. M. and Buck, K. J. (2002) Expression profiling identifies strain-specific changes associated with ethanol withdrawal in mice. Genes Brain and Behavior 1, 35–45.[CrossRef][ISI]

Darstein, M. B., Landwehrmeyer, G. B. and Feuerstein, T. J. (2000) Changes in NMDA receptor subunit gene expression in the rat brain following withdrawal from forced long-term ethanol intake. Naunyn Schmiedeberg's Archives of Pharmacology 361, 206–213.[CrossRef][ISI][Medline]

Davis-Cox, M. I., Fletcher, T. L., Turner, J. N., Szarowski, D. and Shain, W. (1996) Three-day exposure to low-dose ethanol alters guanine nucleotide binding protein expression in the developing rat hippocampus. Journal of Pharmacology and Experimental Therapeutics 276, 758–764.[Abstract]

Davis, K. M. and Wu, J. Y. (2001) Role of glutamatergic and GABAergic systems in alcoholism. Journal of Biomedical Sciences 8, 7–19.[CrossRef]

de Gortari, P., Mendez, M., Rodriguez-Keller, I., Perez-Martinez, L. and Joseph-Bravob, P. (2000) Acute ethanol administration induces changes in TRH and proenkephalin expression in hypothalamic and limbic regions of rat brain. Neurochemistry International 37, 483–496.[CrossRef][ISI][Medline]

Dennis, G. Jr, Sherman, B. T., Hosack, D. A., Yang, J., Gao, W., Lane, H. C. and Lempicki, R. A. (2003) DAVID: Database for annotation, visualization and integrated discovery. Genome Biology 4, P3.

Devaud, L. L., Smith, F. D., Grayson, D. R. and Morrow, A. L. (1995) Chronic ethanol consumption differentially alters the expression of gamma-aminobutyric acidA receptor subunit mRNAs in rat cerebral cortex: competitive, quantitative reverse transcriptase–polymerase chain reaction analysis. Molecular Pharmacology 48, 861–868.[Abstract]

Devaud, L. L., Fritschy, J. M., Sieghart, W. and Morrow, A. L. (1997) Bidirectional alterations of GABA(A) receptor subunit peptide levels in rat cortex during chronic ethanol consumption and withdrawal. Journal of Neurochemistry 69, 126–130.[ISI][Medline]

Devaud, L. L., Fritschy, J. M. and Morrow, A. L. (1998) Influence of gender on chronic ethanol-induced alterations in GABAA receptors in rats. Brain Research 796, 222–230.[CrossRef][ISI][Medline]

Dick, D. M. and Foroud, T. (2003) Candidate genes for alcohol dependence: a review of genetic evidence from human studies. Alcoholism: Clinical and Experimental Research 27, 868–879.[ISI][Medline]

Drago, J., Padungchaichot, P., Accili, D. and Fuchs, S. (1998) Dopamine receptors and dopamine transporter in brain function and addictive behaviors: insights from targeted mouse mutants. Developmental Neuroscience 20, 188–203.[CrossRef][ISI][Medline]

Emanuele, M. A., Tentler, J. J., Kirsteins, L., Emanuele, N. V., Lawrence, A. and Kelley, M. R. (1992) The effect of ‘binge’ ethanol exposure on growth hormone and prolactin gene expression and secretion. Endocrinology 131, 2077–2082.[Abstract]

Eravci, M., Grosspietsch, T., Pinna, G. et al., (1997) Dopamine receptor gene expression in an animal model of ‘behavioral dependence’ on ethanol. Brain Research. Molecular Brain Research 50, 221–229.[ISI][Medline]

Eravci, M., Kley, S., Pinna, G., Prengel, H., Brodel, O., Hiedra, L., Meinhold, H. and Baumgartner, A. (1999) Gene expression of glucose transporters and glycolytic enzymes in the CNS of rats behaviorally dependent on ethanol. Brain Research. Molecular Brain Research 65, 103–111.[ISI][Medline]

Eravci, M., Schulz, O., Grospietsch, T., Pinna, G., Brodel, O., Meinhold, H. and Baumgartner, A. (2000) Gene expression of receptors and enzymes involved in GABAergic and glutamatergic neurotransmission in the CNS of rats behaviourally dependent on ethanol. British Journal of Pharmacology 131, 423–432.[Abstract/Free Full Text]

Eriksson, K. (1968) Genetic selection for voluntary alcohol consumption in the albino rat. Science 159, 739–741.[ISI]

Fan, L., van der Brug, M., Chen, W., Dodd, P. R., Matsumoto, I., Niwa, S. and Wilce, P. A. (1999) Increased expression of mitochondrial genes in human alcoholic brain revealed by differential display. Alcoholism: Clinical and Experimental Research 23, 408–413.[ISI][Medline]

Fan, L., Jaquet, V., Dodd, P. R., Chen, W. and Wilce, P. A. (2001) Molecular cloning and characterization of hNP22: a gene up-regulated in human alcoholic brain. Journal of Neurochemistry 76, 1275–1281.[CrossRef][ISI][Medline]

Follesa, P., Giggio, F., Mancuso, L., Cabras, S., Caria, S., Gorini, G., Manca, A., Orru, A. and Biggio, G. (2004) Ethanol withdrawal-induced up-regulation of the {alpha}2 subunit of the GABAA receptor and its prevention by diazepam or {gamma}-hydroxybutyric acid. Molecular Brain Research 120, 130–137.[ISI][Medline]

Gora-Maslak, G., McClearn, G. E., Crabbe, J. C., Phillips, T. J., Belknap, J. K. and Plomin, R. (1991) Use of recombinant inbred strains to identify quantitative trait loci in psychopharmacology. Psychopharmacology (Berl) 104, 413–424.[ISI][Medline]

Grant, K. A. and Lovinger, D. M. (1995) Cellular and behavioral neurobiology of alcohol: receptor-mediated neuronal processes. Clinical Neuroscience 3, 155–164.[ISI][Medline]

Grummer, M. A. and Zachman, R. D. (1995) Prenatal ethanol consumption alters the expression of cellular retinol binding protein and retinoic acid receptor mRNA in fetal rat embryo and brain. Alcoholism: Clinical and Experimental Research 19, 1376–1381.[ISI][Medline]

Hardy, P. A., Chen, W. and Wilce, P. A. (1999) Chronic ethanol exposure and withdrawal influence NMDA receptor subunit and splice variant mRNA expression in the rat cerebral cortex. Brain Research 819, 33–39.[CrossRef][ISI][Medline]

Herz, A. (1997) Endogenous opioid systems and alcohol addiction. Psychopharmacology (Berl) 129, 99–111.[CrossRef][ISI][Medline]

Hirouchi, M., Hashimoto, T. and Kuriyama, K. (1993) Alteration of GABAA receptor alpha 1-subunit mRNA in mouse brain following continuous ethanol inhalation. European Journal of Pharmacology 247, 127–130.[CrossRef][Medline]

Hu, X. J., Follesa, P. and Ticku, M. K. (1996) Chronic ethanol treatment produces a selective upregulation of the NMDA receptor subunit gene expression in mammalian cultured cortical neurons. Brain Research. Molecular Brain Research 36, 211–218.[ISI][Medline]

Hwang, B. H., Froehlich, J. C., Hwang, W. S., Lumeng, L. and Li, T. K. (1998) More vasopressin mRNA in the paraventricular hypothalamic nucleus of alcohol-preferring rats and high alcohol-drinking rats selectively bred for high alcohol preference. Alcoholism: Clinical and Experimental Research 22, 664–669.[ISI][Medline]

Jamensky, N. T. and Gianoulakis, C. (1999) Comparison of the proopiomelanocortin and proenkephalin opioid peptide systems in brain regions of the alcohol-preferring C57BL/6 and alcohol-avoiding DBA/2 mice. Alcohol 18, 177–187.[CrossRef][ISI][Medline]

Katsura, M., Ohkuma, S., Jun, X., Tsujimura, A. and Kuriyama, K. (1995a) Ethanol stimulates diazepam binding inhibitor (DBI) mRNA expression in primary cultured neurons. Brain Research. Molecular Brain Research 34, 355–359.[ISI][Medline]

Katsura, M., Ohkuma, S., Tsujimura, A. and Kuriyama, K. (1995b) Increase of diazepam binding inhibitor mRNA levels in the brains of chronically ethanol-treated and -withdrawn mice. Journal of Pharmacology and Experimental Therapeutics 273, 1529–1533.[Abstract]

Katsura, M., Ohkuma, S., Tsujimura, A., Xu, J., Hibino, Y., Ishikawa, E. and Kuriyama, K. (1998) Functional involvement of benzodiazepine receptors in ethanol-induced increases of diazepam binding inhibitor (DBI) and its mRNA in the mouse brain. Brain Research Molecular Brain Research 54, 124–132.[ISI][Medline]

Kelley, M. R., Jurgens, J. K., Tentler, J., Emanuele, N. V., Blutt, S. E. and Emanuele, M. A. (1993) Coupled reverse transcription–polymerase chain reaction (RT–PCR) technique is comparative, quantitative, and rapid: uses in alcohol research involving low abundance mRNA species such as hypothalamic LHRH and GRF. Alcohol 10, 185–189.[CrossRef][ISI][Medline]

Kieffer, B. L. and Gaveriaux-Ruff, C. (2002) Exploring the opioid system by gene knockout. Progress in Neurobiology 66, 285–306.[CrossRef][ISI][Medline]

Kim, J. H., Kim, J. E., Kim, H. J., Roh, G. S., Yoo, J. M., Kang, S. S., Cho, Y. Y., Cho, G. J. and Choi, W. S. (2001) Ethanol decreases the expression of mitochondrial cytochrome c oxidase mRNA in the rat. Neuroscience Letters 305, 107–110.[CrossRef][ISI][Medline]

Kinoshita, H., Jessop, D. S., Finn, D. P., Coventry, T. L., Roberts, D. J., Ameno, K., Ijiri, I. and Harbuz, M. S. (2000) Acute ethanol decreases NPY mRNA but not POMC mRNA in the arcuate nucleus. Neuroreport 11, 3517–3519.[ISI][Medline]

Krishnan-Sarin, S., Wand, G. S., Li, X. W., Portoghese, P. S. and Froehlich, J. C. (1998) Effect of mu opioid receptor blockade on alcohol intake in rats bred for high alcohol drinking. Pharmacology Biochemistry and Behavior 59, 627–635.[CrossRef][ISI][Medline]

Kumer, S. C. and Vrana, K. E. (1996) Intricate regulation of tyrosine hydroxylase activity and gene expression. Journal of Neurochemistry 67, 443–462.[ISI][Medline]

Lee, S. and Rivier, C. (1994) Interaction between alcohol and interleukin-1 beta on ACTH secretion and the expression of immediate early genes in the hypothalamus. Molecular and Cellular Neuroscience 5, 442–450.[CrossRef][ISI][Medline]

Lee, S., Smith, G. W., Vale, W., Lee, K. F. and Rivier, C. (2001) Mice that lack corticotropin-releasing factor (CRF) receptors type 1 show a blunted ACTH response to acute alcohol despite up-regulated constitutive hypothalamic CRF gene expression. Alcoholism: Clinical and Experimental Research 25, 427–433.[CrossRef][ISI][Medline]

Lewohl, J. M., Wang, L., Miles, M. F., Zhang, L., Dodd, P. R. and Harris, R. A. (2000) Gene expression in human alcoholism: microarray analysis of frontal cortex. Alcoholism: Clinical and Experimental Research 24, 1873–1882.[ISI][Medline]

Li, X. W., Li, T. K. and Froehlich, J. C. (1998) Enhanced sensitivity of the nucleus accumbens proenkephalin system to alcohol in rats selectively bred for alcohol preference. Brain Research 794, 35–47.[CrossRef][ISI][Medline]

Lingford-Hughes, A. R., Davies, S. J., McIver, S., Williams, T. M., Daglish, M. R. and Nutt, D. J. (2003) Addiction. British Medical Bulletin 65, 209–222.[Abstract/Free Full Text]

Lintunen, M., Raatesalmi, K., Sallmen, T., Anichtchik, O., Karlstedt, K., Kaslin, J., Kiianmaa, K., Korpi, E. R. and Panula, P. (2002) Low brain histamine content affects ethanol-induced motor impairment. Neurobiology of Disease 9, 94–105.[CrossRef][ISI][Medline]

Lumeng, L., Hawkins, T. D. and Li, T. K. (1977) New strains of rats with alcohol preference and non-preference. In Alcohol and aldehyde metabolizing systems, Thurman, R. G., Williamson, J. R., Drott, H. and Chance, B. eds, pp. 537–544. Academic Press, New York.

Maciejewski-Lenoir, D. (1993) Chronic prenatal ethanol exposure does not affect the expression of selected genes in rat brain development. Alcohol and Alcoholism 28, 401–412.[Abstract]

MacLennan, A. J., Lee, N. and Walker, D. W. (1995) Chronic ethanol administration decreases brain-derived neurotrophic factor gene expression in the rat hippocampus. Neuroscience Letters 197, 105–108.[CrossRef][ISI][Medline]

Mahmoudi, M., Kang, M. H., Tillakaratne, N., Tobin, A. J. and Olsen, R. W. (1997) Chronic intermittent ethanol treatment in rats increases GABA(A) receptor alpha4-subunit expression: possible relevance to alcohol dependence. Journal of Neurochemistry 68, 2485–2492.[ISI][Medline]

Maier, S. E., Cramer, J. A., West, J. R. and Sohrabji, F. (1999) Alcohol exposure during the first two trimesters equivalent alters granule cell number and neurotrophin expression in the developing rat olfactory bulb. Journal of Neurobiology 41, 414–423.[CrossRef][ISI][Medline]

Marinelli, P. W., Kiianmaa, K. and Gianoulakis, C. (2000) Opioid propeptide mRNA content and receptor density in the brains of AA and ANA rats. Life Sciences 66, 1915–1927.[CrossRef][ISI][Medline]

Matthews, D. B., Devaud, L. L., Fritschy, J. M., Sieghart, W. and Morrow, A. L. (1998) Differential regulation of GABA(A) receptor gene expression by ethanol in the rat hippocampus versus cerebral cortex. Journal of Neurochemistry 70, 1160–1166.[ISI][Medline]

Mayfield, R. D., Lewohl, J. M., Dodd, P. R., Herlihy, A., Liu, J. and Harris, R. A. (2002) Patterns of gene expression are altered in the frontal and motor cortices of human alcoholics. Journal of Neurochemistry 81, 802–813.[CrossRef][ISI][Medline]

McBride, W. J. and Li, T. K. (1998) Animal models of alcoholism: neurobiology of high alcohol-drinking behavior in rodents. Critical Reviews in Neurobiology 12, 339–369.[ISI][Medline]

McClearn, G. E. and Rodgers, D. A. (1959) Differences in alcohol preference among inbred strains of mice. Quarterly Journal of the Study of Alcohol 20, 691–695.

McClearn, G. E. and Kikihana, R. (1981) Selective breeding for ethanol sensitivity: short-sleep and long-sleep mice. In Development of Animal Models as Pharmacogenetic Tools, McClearn, G. E., Deitrich, R. A., and Erwin, V. G. eds, pp. 81–113, Government Printing Office, Washington, DC.

McGue, M., Slutske, W. and Iacono, W. G. (1999) Personality and substance use disorders: II. Alcoholism versus drug use disorders. Journal of Consulting and Clinical Psychology 67, 394–404.[CrossRef][ISI][Medline]

McKenna, M. C., Stevenson, J. H., Huang, X. and Hopkins, I. B. (2000) Differential distribution of the enzymes glutamate dehydrogenase and aspartate aminotransferase in cortical synaptic mitochondria contributes to metabolic compartmentation in cortical synaptic terminals. Neurochemistry International 37, 229–241.[ISI][Medline]

Meldrum, B. S. (2000) Glutamate as a neurotransmitter in the brain: review of physiology and pathology. Journal of Nutrition 130, 1007–1015S.[Abstract/Free Full Text]

Mhatre, M. C. and Ticku, M. K. (1992) Chronic ethanol administration alters gamma-aminobutyric acidA receptor gene expression. Molecular Pharmacology 42, 415–422.[Abstract]

Miotto, K., Kauffman, D., Anton, B., Kieth, D. E. Jr and Evans, C. J. (1996) Human opioid receptors: chromosomal mapping and mRNA localization. National Institute on Drug Abuse Research Monograph 161, 72–82.

Mitchell, S. E. and Snyder-Keller, A. (2003) c-fos and cleaved caspase-3 expression after perinatal exposure to ethanol, cocaine, or the combination of both drugs. Developmental Brain Research 147, 107–117.[ISI][Medline]

Mitsuyama, H., Little, K. Y., Sieghart, W., Devaud, L. L. and Morrow, A. L. (1998) GABA(A) receptor alpha1, alpha4, and beta3 subunit mRNA and protein expression in the frontal cortex of human alcoholics. Alcoholism: Clinical and Experimental Research 22, 815–822.[ISI][Medline]

Mochly-Rosen, D., Chang, F. H., Cheever, L., Kim, M., Diamond, I. and Gordon, A. S. (1988) Chronic ethanol causes heterologous desensitization of receptors by reducing alpha s messenger RNA. Nature 333, 848–850.[CrossRef][ISI][Medline]

Montpied, P., Morrow, A. L., Karanian, J. W., Ginns, E. I., Martin, B. M. and Paul, S. M. (1991) Prolonged ethanol inhalation decreases gamma-aminobutyric acidA receptor alpha subunit mRNAs in the rat cerebral cortex. Molecular Pharmacology 39, 157–163.[Abstract]

Moore, D. B., Walker, D. W. and Heaton, M. B. (1999) Neonatal ethanol exposure alters bcl-2 family mRNA levels in the rat cerebellar vermis. Alcoholism: Clinical and Experimental Research 23, 1251–1261.[ISI][Medline]

Naassila, M. and Daoust, M. (2002) Effect of prenatal and postnatal ethanol exposure on the developmental profile of mRNAs encoding NMDA receptor subunits in rat hippocampus. Journal of Neurochemistry 80, 850–860.[CrossRef][ISI][Medline]

Nagahara, A. H. and Handa, R. J. (1995) Fetal alcohol exposure alters the induction of immediate early gene mRNA in the rat prefrontal cortex after an alternation task. Alcoholism: Clinical and Experimental Research 19, 1389–1397.[ISI][Medline]

Nevo, I., Langlois, X., Laporte, A. M., Kleven, M., Koek, W., Lima, L., Maudhuit, C., Martres, M. P. and Hamon, M. (1995) Chronic alcoholization alters the expression of 5-HT1A and 5-HT1B receptor subtypes in rat brain. European Journal of Pharmacology 281, 229–239.[CrossRef][ISI][Medline]

Nixon, K., Hughes, P. D., Amsel, A. and Leslie, S. W. (2004) NMDA receptor subunit expression after combined prenatal and postnatal exposure to ethanol. Alcoholism: Clinical and Experimental Research 28, 105–112.[ISI][Medline]

Ogilvie, K., Lee, S. and Rivier, C. (1997a) Effect of three different modes of alcohol administration on the activity of the rat hypothalamic–pituitary–adrenal axis. Alcoholism: Clinical and Experimental Research 21, 467–476.[ISI][Medline]

Ogilvie, K. M., Lee, S. and Rivier, C. (1997b) Role of arginine vasopressin and corticotropin-releasing factor in mediating alcohol-induced adrenocorticotropin and vasopressin secretion in male rats bearing lesions of the paraventricular nuclei. Brain Research 744, 83–95.[CrossRef][ISI][Medline]

Ohkuma, S., Katsura, M. and Tsujimura, A. (2001) Alterations in cerebral diazepam binding inhibitor expression in drug dependence: a possible biochemical alteration common to drug dependence. Life Sciences 68, 1215–1222.[CrossRef][ISI][Medline]

Pandey, S. C., Mittal, N., Lumeng, L. and Li, T. K. (1999) Involvement of the cyclic AMP-responsive element binding protein gene transcription factor in genetic preference for alcohol drinking behavior. Alcoholism: Clinical and Experimental Research 23, 1425–1434.[ISI][Medline]

Paula-Barbosa, M. M., Silva, S. M., Andrade, J. P., Cadete-Leite, A. and Madeira, M. D. (2001) Nerve growth factor restores mRNA levels and the expression of neuropeptides in the suprachiasmatic nucleus of rats submitted to chronic ethanol treatment and withdrawal. Journal of Neurocytology 30, 195–207.[CrossRef][ISI][Medline]

Petrie, J., Sapp, D. W., Tyndale, R. F., Park, M. K., Fanselow, M. and Olsen, R. W. (2001) Altered gabaa receptor subunit and splice variant expression in rats treated with chronic intermittent ethanol. Alcoholism: Clinical and Experimental Research 25, 819–828.[CrossRef][ISI][Medline]

Plotkin, S. R., Banks, W. A., Waguespack, P. J. and Kastin, A. J. (1997) Ethanol alters the concentration of Met-enkephalin in brain by affecting peptide transport system-1 independent of preproenkephalin mRNA. Journal of Neuroscience Research 48, 273–280.[CrossRef][ISI][Medline]

Pompei, P., Angeletti, S., Ciccocioppo, R., Colombo, G., Gessa, G. L. and Massi, M. (1998) Preprotachykinin-A gene expression in the forebrain of Sardinian alcohol-preferring and -nonpreferring rats. Brain Research. Molecular Brain Research 56, 277–280.[ISI][Medline]

Putzke, J., De Beun, R., Schreiber, R., De Vry, J., Tolle, T. R., Zieglgansberger, W. and Spanagel, R. (1998) Long-term alcohol self-administration and alcohol withdrawal differentially modulate microtubule-associated protein 2 (MAP2) gene expression in the rat brain. Brain Research. Molecular Brain Research 62, 196–205.[ISI][Medline]

Reich, T., Edenberg, H. J., Goate, A. et al., (1998) Genome-wide search for genes affecting the risk for alcohol dependence. American Journal of Medical Genetics 81, 207–215.[CrossRef][ISI][Medline]

Reilly, M. T. and Buck, K. J. (2000) GABA(A) receptor beta(2) subunit mRNA content is differentially regulated in ethanol-dependent DBA/2J and C57BL/6J mice. Neurochemistry International 37, 443–452.[CrossRef][ISI][Medline]

Ren, L. Q., Garrett, D. K., Syapin, M. and Syapin, P. J. (2000) Differential fibronectin expression in activated C6 glial cells treated with ethanol. Molecular Pharmacology 58, 1303–1309.[ISI][Medline]

Rimondini, R., Arlinde, C., Sommer, W. and Heilig, M. (2002) Long-lasting increase in voluntary ethanol consumption and transcriptional regulation in the rat brain after intermittent exposure to alcohol. The FASEB Journal 16, 27–35.[Abstract/Free Full Text]

Rivier, C. and Lee, S. (1996) Acute alcohol administration stimulates the activity of hypothalamic neurons that express corticotropin-releasing factor and vasopressin. Brain Research 726, 1–10.[CrossRef][ISI][Medline]

Ryabinin, A. E. and Wang, Y. M. (1998) Repeated alcohol administration differentially affects c-Fos and FosB protein immunoreactivity in DBA/2J mice. Alcoholism: Clinical and Experimental Research 22, 1646–1654.[ISI][Medline]

Ryabinin, A. E., Wang, Y. M., Freeman, P. and Risinger, F. O. (1999) Selective effects of alcohol drinking on restraint-induced expression of immediate early genes in mouse brain. Alcoholism: Clinical and Experimental Research 23, 1272–1280.[ISI][Medline]

Saito, M., Smiley, J., Toth, R. and Vadasz, C. (2002) Microarray analysis of gene expression in rat hippocampus after chronic ethanol treatment. Neurochemistry Research 27, 1221–1229.[CrossRef][ISI][Medline]

Sall, J. M., Morehead, M., Murphy, S., Goldman, H. and Walker, P. D. (1996) Alterations in CNS gene expression in a rodent model of moderate traumatic brain injury complicated by acute alcohol intoxication. Experimental Neurology 139, 257–268.[CrossRef][ISI][Medline]

Scanlon, M. N., Lazar-Wesley, E., Grant, K. A. and Kunos, G. (1992) Proopiomelanocortin messenger RNA is decreased in the mediobasal hypothalamus of rats made dependent on ethanol. Alcoholism: Clinical and Experimental Research 16, 1147–1151.[ISI][Medline]

Schafer, G. L., Crabbe, J. C. and Wiren, K. M. (2001) Ethanol-regulated gene expression of neuroendocrine specific protein in mice: brain region and genotype specificity. Brain Research 897, 139–149.[CrossRef][ISI][Medline]

Scott, H. C., Zoeller, R. T. and Rudeen, P. K. (1995) Acute prenatal ethanol exposure and luteinizing hormone-releasing hormone messenger RNA expression in the fetal mouse brain. Alcoholism: Clinical and Experimental Research 19, 153–159.[ISI][Medline]

Silva, S. M., Madeira, M. D., Ruela, C. and Paula-Barbosa, M. M. (2002) Prolonged alcohol intake leads to irreversible loss of vasopressin and oxytocin neurons in the paraventricular nucleus of the hypothalamus. Brain Research 925, 76–88.[CrossRef][ISI][Medline]

Singh, L. D., Singh, S. P., Handa, R. K., Ehmann, S. and Snyder, A. K. (1996a) Effects of ethanol on GLUT1 protein and gene expression in rat astrocytes. Metabolic Brain Disease 11, 343–357.[ISI][Medline]

Singh, S. P., Ehmann, S. and Snyder, A. K. (1996b) Ethanol-induced changes in insulin-like growth factors and IGF gene expression in the fetal brain. Proceedings of the Society for Experimental Biology and Medicine 212, 349–354.[Abstract]

Sohma, H., Hashimoto, E., Shirasaka, T., Tsunematsu, R., Ozawa, H., Boissl, K. W., Boning, J., Riederer, P. and Saito, T. (1999) Quantitative reduction of type I adenylyl cyclase in human alcoholics. Biochimica et Biophysica Acta 1454, 11–18.[ISI][Medline]

Sokolov, B. P., Jiang, L., Trivedi, N. S. and Aston, C. (2003) Transcription profiling reveals mitochondrial, ubiquitin and signaling systems abnormalities in postmortem brains from subjects with a history of alcohol abuse or dependence. Journal of Neuroscience Research 72, 756–767.[CrossRef][ISI][Medline]

Soszynski, P. A. and Frohman, L. A. (1992) Inhibitory effects of ethanol on the growth hormone (GH)-releasing hormone-GH-insulin-like growth factor-I axis in the rat. Endocrinology 131, 2603–2608.[Abstract]

Sterling, P. and Eyer, J. (1988) Allostasis: a new paradigm to explain arousal pathology. In Handbook of Life Stress, Cognition, and Health, Fisher, S., Reason, J., eds, Wiley, New York.

Suzuki, R., Lumeng, L., McBride, W. J., Li, T.-K. and Hwang, B. H. (2004) Reduced neuropeptide Y mRNA expression in the central nucleus of amygdala of alcohol preferring (P) rats: its potential involvement in alcohol preference and anxiety. Brain Research 1014, 251–254.[CrossRef][ISI][Medline]

Svenneby, G. and Torgner, I. A. (1987) Localization and function of glutamine synthetase and glutaminase. Biochemical Society Transactions 15, 213–215.[ISI][Medline]

Szot, P., White, S. S., Veith, R. C. and Rasmussen, D. D. (1999) Reduced gene expression for dopamine biosynthesis and transport in midbrain neurons of adult male rats exposed prenatally to ethanol. Alcoholism: Clinical and Experimental Research 23, 1643–1649.[ISI][Medline]

Tajuddin, N. F. and Druse, M. J. (1998) Effects of chronic ethanol consumption and aging on proenkephalin and neurotensin. Alcoholism: Clinical and Experimental Research 22, 1152–1160.[ISI][Medline]

Thibault, C., Lai, C., Wilke, N. et al., (2000) Expression profiling of neural cells reveals specific patterns of ethanol-responsive gene expression. Molecular Pharmacology 58, 1593–1600.[ISI][Medline]

Toropainen, M., Nakki, R., Honkanen, A., Rosenberg, P. H., Laurie, D. J., Pelto-Huikko, M., Koistinaho, J., Eriksson, C. J. and Korpi, E. R. (1997) Behavioral sensitivity and ethanol potentiation of the N-methyl-D-aspartate receptor antagonist MK-801 in a rat line selected for high ethanol sensitivity. Alcoholism: Clinical and Experimental Research 21, 666–671.[ISI][Medline]

Treadwell, J. A. and Singh, S. M. (2004) Microarray analysis of mouse brain gene expression following acute ethanol treatment. Neurochemical Research 29, 357–369.[CrossRef][ISI][Medline]

Tunici, P., Schiaffonati, L., Rabellotti, E., Tiberio, L., Perin, A. and Sessa, A. (1999) In vivo modulation of 73 kDa heat shock cognate and 78 kDa glucose-regulating protein gene expression in rat liver and brain by ethanol. Alcoholism: Clinical and Experimental Research 23, 1861–1867.[ISI][Medline]

Valles, S., Pitarch, J., Renau-Piqueras, J. and Guerri, C. (1997) Ethanol exposure affects glial fibrillary acidic protein gene expression and transcription during rat brain development. Journal of Neurochemistry 69, 2484–2493.[ISI][Medline]

Wilson, M. E., Marshall, M. T., Bollnow, M. R., McGivern, R. F. and Handa, R. J. (1995) Gonadotropin-releasing hormone mRNA and gonadotropin beta-subunit mRNA expression in the adult female rat exposed to ethanol in utero. Alcoholism: Clinical and Experimental Research 19, 1211–1218.[ISI][Medline]

Winkler, A. and Spanagel, R. (1998) Differences in the kappa opioid receptor mRNA content in distinct brain regions of two inbred mice strains. Neuroreport 9, 1459–1464.[ISI][Medline]

Winkler, A., Buzas, B., Siems, W. E., Heder, G. and Cox, B. M. (1998a) Effect of ethanol drinking on the gene expression of opioid receptors, enkephalinase, and angiotensin-converting enzyme in two inbred mice strains. Alcoholism: Clinical and Experimental Research 22, 1262–1271.[ISI][Medline]

Winkler, A., Rottmann, M., Heder, G., Hyytia, P., Siems, W. E. and Melzig, M. F. (1998b) Gene expression and activity of specific opioid-degrading enzymes in different brain regions of the AA and ANA lines of rats. Biochimica et Biophysica Acta 1406, 219–227.[ISI][Medline]

Winkler, A., Mahal, B., Kiianmaa, K., Zieglgansberger, W. and Spanagel, R. (1999) Effects of chronic alcohol consumption on the expression of different NR1 splice variants in the brain of AA and ANA lines of rats. Brain Research. Molecular Brain Research 72, 166–175.[ISI][Medline]

Xia, J., Simonyi, A. and Sun, G. Y. (1998) Changes in IP3R1 and SERCA2b mRNA levels in the gerbil brain after chronic ethanol administration and transient cerebral ischemia–reperfusion. Brain Research. Molecular Brain Research 56, 22–28.[ISI][Medline]

Xu, Y., Ehringer, M., Yang, F. and Sikela, J. M. (2001) Comparison of global brain gene expression profiles between inbred long-sleep and inbred short-sleep mice by high-density gene array hybridization. Alcoholism: Clinical and Experimental Research 25, 810–818.[CrossRef][ISI][Medline]

Yagle, K. and Costa, L. G. (1999) Effects of alcohol on immediate-early gene expression in primary cultures of rat cortical astrocytes. Alcoholism: Clinical and Experimental Research 23, 446–455.[ISI][Medline]

Zhou, Y., Franck, J., Spangler, R., Maggos, C. E., Ho, A. and Kreek, M. J. (2000) Reduced hypothalamic POMC and anterior pituitary CRF1 receptor mRNA levels after acute, but not chronic, daily ‘binge’ intragastric alcohol administration. Alcoholism: Clinical and Experimental Research 24, 1575–1582.[CrossRef][ISI][Medline]

Zoeller, R. T., Butnariu, O. V., Fletcher, D. L. and Riley, E. P. (1994) Limited postnatal ethanol exposure permanently alters the expression of mRNAS encoding myelin basic protein and myelin-associated glycoprotein in cerebellum. Alcoholism: Clinical and Experimental Research 18, 909–916.[ISI][Medline]





This Article
Abstract
Full Text (PDF)
All Versions of this Article:
40/1/63    most recent
agh119v1
Alert me when this article is cited
Alert me if a correction is posted
Services
Email this article to a friend
Similar articles in this journal
Similar articles in ISI Web of Science
Similar articles in PubMed
Alert me to new issues of the journal
Add to My Personal Archive
Download to citation manager
Search for citing articles in:
ISI Web of Science (2)
Request Permissions
Google Scholar
Articles by WORST, T. J.
Articles by VRANA, K. E.
PubMed
PubMed Citation
Articles by WORST, T. J.
Articles by VRANA, K. E.