13. Dopaminergic compounds - preclinical data

Editors: Spanagel, Rainer; Mann, Karl F.

Title: Drugs for Relapse Prevention of Alcoholism, 1st Edition

Copyright 2005 Springer

> Table of Contents > Dopaminergic compounds: preclinical data

Dopaminergic compounds: preclinical data

Friedbert Weiss

Department of Neuropharmacology, The Scripps Research Institute (CVN 15), 10550 North Torrey Pines Road, La Jolla, CA 92037, USA

Introduction

Like other addictive disorders, alcoholism is characterized by long-lasting vulnerability to relapse after cessation of drinking. Increasing attention has, therefore, been directed toward identifying specific risk factors for relapse and to establish the neurobiological mechanisms by which these factors convey vulnerability to relapse. At the same time, relapse prevention has emerged as an important focus of treatment and medication development efforts.

An important neuropharmacological substrate for many of the neurobehavioral effects of ethanol is the mesocorticolimbic dopamine (DA) system. Evidence has accumulated in recent years implicating this system also in a range of behavioral and neurobiological processes relevant for alcohol craving and relapse. These findings identify DA transmission as a possible pharmacotherapeutic target for relapse prevention.

Evidence implicating dopamine neurotransmission as a target for relapse prevention

Craving and relapse associated with ethanol cue exposure

Ethanol acutely increases the activity of the mesolimbic DA reward pathway [1, 2, 3, 4 and 5] and this effect has been widely implicated as a mechanism by which ethanol exerts its reinforcing actions [6, 7, 8 and 9]. However, not only ethanol consumption [5, 10], but also exposure to environmental stimuli associated with ethanol can activate DA transmission in the nucleus accumbens [5, 10, 11 and 12]. Thus, in addition to its role in the reinforcing effects of ethanol, mesolimbic DA transmission may mediate the incentive-motivational effects of ethanol-related environmental stimuli that are thought to be relevant for ethanol-seeking, craving and relapse. The conditioning of ethanol's pharmacological actions with discrete environmental stimuli is a major factor in the abuse potential of this drug [13]. Ethanol cues can evoke drug desire that may lead to the resumption of drinking in abstinent alcoholics [14, 15, 16, 17, 18, 19, 20 and 21]. Drug-related

P.136


stimuli may perhaps also elicit automatic responses that lead to drug-seeking behavior and relapse without the intervention of distinct feelings of craving [22, 23]. A role of DA in such conditioned responses to ethanol cues would be consistent with the established general role of mesocorticolimbic DA transmission in incentive learning and conditioned reinforcement associated with both natural and drug rewards (e.g., [24, 25 and 26]). In support of this possibility, alcohol-related stimuli not only increase accumbal DA release in rats [5, 10, 11 and 12], but activate the ventral striatum (i.e., a DA-rich brain region) in abstinent alcoholics [27]. Consistent with this hypothesis as well, pharmacological activation of DA transmission in the nucleus accumbens enhances the ability of ethanol-paired stimuli to function as conditioned reinforcers [28].

Craving elicited by priming doses of ethanol

It is well established that small doses of drugs of abuse including ethanol, rather than reducing drug desire, elicit further drug craving (e.g., [29, 30]). Moreover, in alcoholics, the first drink after abstinence is often associated with loss of control leading to severe intoxication and return to continued alcohol abuse [30]. A role for DA in this aspect of relapse has emerged in controlled laboratory studies showing that the D2-preferring DA receptor antagonist haloperidol reduces alcohol craving in response to a priming dose of ethanol in alcoholic patients [31]. Haloperidol also blocked the stimulant and euphorigenic effects of ethanol in social drinkers [32]. The attenuation of ethanol's priming effects supports the view that craving and loss of control following resumption of drinking are linked to DA-activating effects of ethanol. More generally, these findings support the wealth of preclinical evidence that has implicated DA as a major neuropharmacological substrate for ethanol reward.

Neuroadaptation induced by chronic ethanol

In contrast to the acute effects of ethanol that activate mesolimbic DA neurotransmission, chronic dependence-inducing ethanol treatments lead to the development of DA hypofunction. Animals chronically exposed to ethanol show impaired DA synthesis (e.g., [33, 34 and 35]) paired with elevated DA transporter levels and, thus, presumably enhanced clearance of synaptic DA [34]. Impairments in mesolimbic DA function are particularly evident during withdrawal with marked decrements in the activity of ventral tegmental DA neurons [36, 37] and reduced extracellular DA levels in the nucleus accumbens [38, 39]. Renewed ethanol administration reverses these functional deficits as well as behavioral signs of withdrawal in rats [39, 40]. Consistent with these findings in animals, the DA D2 agonist bromocriptine ameliorates ethanol

P.137


withdrawal symptoms, including anxiety, depression, and restlessness in alcoholics [41]. Deficient DA function, therefore, is likely to represent a neural basis for the state of dysphoria and negative affect that accompanies ethanol withdrawal such that restoration of normal DA function by appropriate pharmacological agents may be a suitable strategy for relapse prevention.

Deficits in DA neurotransmission induced by chronic ethanol appear to be long lasting. Midbrain DA neural activity still shows suppression three days after withdrawal [42]. Measures of accumbal DA synthesis and turnover indicate that the release of DA is deficient as late as two months after withdrawal [43], and impairments in DA D1-mediated signal transduction have been observed as late as four months following withdrawal from long-term ethanol exposure [44]. The persistent impairment in the functioning of mesolimbic DA transmission likely has implications for vulnerability to relapse during protracted ethanol withdrawal. Indeed, clinical studies suggest that retardation in the recovery of impairments in DA receptor function or DA-dependent signal transduction is associated with the risk of early relapse [45, 46]. Relapse risk linked to delayed recovery of DA receptor function may not, however, extend to heightened ethanol cue reactivity or cue-induced craving, since a more recent study failed to demonstrate a relationship between DA receptor sensitivity and craving induced by the smell of alcoholic beverages [47]. A role of deficient DA transmission in the resumption of drinking has also found important indirect support by findings that low endogenous DA function is linked to predisposition for increased alcohol preference and intake [48, 49 and 50], whereas pharmacological enhancement of DA synthesis, inhibition of DA degradation, or overexpression of D2 receptors by adenoviral DRD2 gene delivery exert protective effects reducing ethanol preference and intake [58, 51].

Exacerbation of drinking following ethanol deprivation

A well-described phenomenon in the alcohol literature is a marked increase in ethanol consumption that follows periods of alcohol deprivation [52, 53]. This alcohol deprivation effect (ADE) is considered a measure of motivation for alcohol [52, 54], loss of control [55], or relapse [50, 56]. Similarities exist between the alcohol deprivation effect in animals and human alcohol abuse such as enhanced ethanol consumption after abstinence in social drinkers [57], and aspects of the loss of control phenomenon surrounding the first drink after abstinence in alcoholics [30, 58, 59]. With repeated deprivation and increased length of deprivation periods, this phenomenon becomes resistant to manipulations of ethanol concentration, taste, and environmental factors [55, 60, 61]. Moreover, increases in ethanol intake produced by repeated deprivation outlast long abstinence phases [60] and may become irreversible [55]. The ADE, therefore, provides a model to study compulsive alcohol-seeking behavior and loss of control following resumption of drinking.

P.138


Neurochemical studies implicate dysregulation of mesolimbic DA transmission as a factor contributing to enhanced drinking following alcohol deprivation. DA release in the nucleus accumbens as measured by microdialysis [62] as well as electrically evoked release of [3H]DA from accumbal tissue [63] was increased after 2-3 weeks of ethanol deprivation in alcohol-preferring P and Wistar rats following several weeks of free access to ethanol. In addition, under these conditions the ADE was associated with a reduction of autoregulatory increases on accumbal DA overflow induced by D2 receptor stimulation [62]. The latter finding suggests that increased DA activity resulting from compromised D2 autoreceptor function may play a role in the ADE [62], although this interpretation is difficult to reconcile with the wealth of data showing that increased DA function is associated with decreased ethanol preference and intake whereas preference and intake are increased in animals with low DA function (e.g., [48, 49, 50 and 51]). Data from a model that utilized repeated deprivation to induce high ethanol drinking also implicate impaired D1-mediated signal transduction as reflected by a decrease in the efficacy of DA to stimulate striatal adenylyl cyclase activity in deprivation-induced exacerbation of ethanol intake [44]. Thus, persistent ethanol-induced changes in DA receptor function may be responsible for increased ethanol drinking associated with the ADE.

In summary, the literature points toward a prominent role for mesolimbic DA transmission in behavioral and neurobiological processes relevant for alcohol relapse risk. Despite this evidence, only few preclinical studies have explicitly tested the effects of pharmacological manipulation of DA neurotransmission on relapse. The following review will therefore extend beyond data from strict animal models of relapse to include a wider spectrum of experimental approaches relevant for evaluating the potential of DA transmission as a treatment target for relapse prevention. Findings covered by this review are summarized in Figure 1.

Manipulation of dopamine receptors and ethanol-seeking behavior

Conditioned ethanol-seeking behavior

A dopaminergic role in drug-related learning and the potential of DA antagonists to reverse ethanol-seeking induced by alcohol cues has been studied in several animal models. These models have in common that they measure behavior controlled by the incentive-motivational or conditioned reinforcing effects of alcohol-associated environmental stimuli. These models differ, however, in terms of whether or not abstinence or extinction of ethanol-reinforced behavior is imposed before testing for conditioned ethanol-seeking, and whether or not conditioning occurs in conjunction with involuntary or self-administration of ethanol.

P.139


Figure 1. Conditioned reinstatement by olfactory discriminative stimuli associated with ethanol availability (S+) versus nonreward (S-), and modification of this effect by the D1 antagonist SCH23390 and the D2 antagonist eticlopride. Rats were treated with DA antagonists 30 min before reinstatement tests. Panels A, C: Total number of responses. For comparison, the figures show also the mean ( SEM) number of responses during ethanol self-administration (??) and nonreward (??) conditioning sessions collapsed across the last three days of the conditioning phase (SA), as well as the mean ( SEM) number of responses across the last three days of the extinction phase (EXT). * p < 0.05, ** p < 0.01 different from extinction baseline; + p < 0.05, different from vehicle. Panels B, D: Latency to initiate responding. * p < 0.05, ** p < 0.01 different from vehicle. + p < 0.05, ++ p < 0.01 different from 5 g/kg. p < 0.01 different from 10 g/kg. Modified from Ref. 54 (with permission)

Conditioned reinstatement

In the context of the drug addiction literature, conditioned reinstatement refers to the resumption of extinguished instrumental responding induced by noncontingent exposure to a drug-related cue (for review see [64, 65]). Ethanol-associated contextual stimuli reliably elicit recovery of responding at a previously ethanol-paired lever following extinction without further alcohol availability [66, 67, 68, 69, 70, 71 and 72]. Moreover, the behavioral effects of these stimuli are remarkably

P.140


resistant to extinction in that recovery of ethanol-seeking does not diminish when these cues are presented repeatedly under non-reinforced conditions [73].

Reinstatement of ethanol-seeking by ethanol-associated contextual stimuli is sensitive to antagonism of DA transmission [74]. Rats operantly self-administering 10% ethanol in daily 30-min sessions were trained to associate distinct olfactory discriminative stimuli with the availability of ethanol or absence of reward. Following subsequent extinction of ethanol-reinforced responding the animals were exposed to the discriminative stimuli previously predictive of ethanol or no reward. Presentation of the ethanol-associated stimulus reinstated responding at the previously active lever. Both the DA D1-selective antagonist SCH 23390, and the DA D2-selective antagonist eticlopride, dose-dependently attenuated this effect by increasing the latency to initiate responding and decreasing the number of responses (Fig. 1). Similar effects were obtained in rats with a history of ethanol dependence, tested three weeks following withdrawal from a 12-day ethanol vapor inhalation procedure. Interestingly, the dose-effect curve for SCH 23390 and eticlopride in these animals was shifted to the left, resulting in an overall greater inhibition of reinstatement responses at lower doses of these agents (Fig. 2). These findings suggest that in both non-dependent and previously dependent rats, blockade of D1 or D2 receptors attenuates the conditioned incentive effects of ethanol-related contextual stimuli. A likely explanation for the increased DA antagonist potency in previously dependent rats involves the DA hypoactivity produced by chronic ethanol as discussed above. Specifically, reduced synaptic availability of DA at the time of testing may have been a factor contributing to the increased DA antagonist potency in rats with a history of ethanol dependence.

Appetitive ethanol-seeking behavior

Recently, an ethanol self-administration model has been developed [75, 76 and 77] that dissociates ethanol-reinforced consummatory behavior (i.e., ethanol drinking) from appetitive ethanol-seeking responses (i.e., behavior induced and maintained by the incentive-motivational effects of ethanol-associated contextual cues present in the self-administration environment). In this procedure, rats must complete a set of responses at a lever during which time ethanol is not available (appetitive phase). Completion of a given response requirement within a specified time results in retraction of the lever and presentation of a sipper tube containing 10% ethanol from which rats are then allowed to freely drink (consummatory phase). Although this procedure is not, strictly speaking, a model of relapse because no significant period of abstinence is imposed before tests, responding during the appetitive phase provides a measure of the day-to-day strength of animals' motivation to initiate and engage in ethanol-seeking behavior when exposed to the ethanol-predictive stimulus environment of the operant conditioning chamber. Thus, pharmacological

P.141


interventions that attenuate appetitive responding in this model may provide relevant leads as to potential treatment targets for conditioned ethanol-seeking and craving.

Figure 2. Dose-effect curves for the attenuation of conditioned reinstatement by SCH 23390 and eticlopride in nondependent rats and rats with a history of prior ethanol dependence. Tests were conducted in nondependent rats prior to a 12-day ethanol vapor inhalation procedure and, again, beginning on day 21 after termination of ethanol vapor exposure (Postdependent). Drug effects are expressed as mean ( SEM) percent inhibition of S+-induced responding compared to the effects of vehicle injection. * p < 0.05 ** p < 0.01 different from Nondependent. Reproduced from Ref. 51 (with permission).

Data generated with this model have revealed that appetitively motivated responding preceding the availability of ethanol is more sensitive to reversal by dopamine D2-selective antagonists than actual ethanol intake [78, 79].

P.142


Raclopride, microinjected into the nucleus accumbens, delayed the onset of appetitive responding and decreased the total number of responses. In contrast, only a single high dose of raclopride reduced the consummatory response. Systemic administration of another D2 antagonist, remoxipride, produced even more selective effects on appetitive ethanol-seeking by dose-dependently decreasing the number of appetitive responses, while having no effect on ethanol consumption.

The findings above demonstrate that both conditioned reinstatement and appetitive ethanol-seeking behavior are sensitive to reversal by DA antagonists. A role of DA in cue-controlled ethanol-seeking is supported further by a study that examined the effects of the D2 antagonist haloperidol on responding maintained by a conditioned stimulus (CS) that had been discretely paired with ethanol-reinforced responses [80]. Haloperidol significantly reduced conditioned responding at a previously active lever. Only a single dose was tested. However, the effects of this dose were selective for responding at the lever paired with the ethanol cue and responding at a control lever remained unaltered. This observation, therefore, lends additional support for the conclusion that DA transmission participates in mediating ethanol-seeking behavior initiated and maintained by ethanol-related environmental stimuli.

Expression of conditioned place preference

Place conditioning procedures permit examination of neuropharmacological substrates involved in the acquisition and expression of the conditioned reinforcing effects of ethanol. In this model, manipulations that interfere specifically with the expression of conditioned place preference (CPP), once acquired, provide information on the neuropharmacological basis of conditioned ethanol-seeking behavior, whereas interference with the acquisition of CPP is relevant for the understanding of mechanisms mediating the acute reinforcing effects of ethanol or the learning of Pavlovian associations. Like the model of appetitive ethanol-seeking behavior, studies on the expression of ethanol CPP have not involved imposition of abstinence and, thus, have limitations with respect to providing a measure of relapse. Nonetheless, the degree of preference for a previously ethanol-paired environment provides an index of the strength of ethanol-seeking associated with the incentive-motivational effects of an alcohol-associated stimulus context.

The strongest evidence in favor of a dopaminergic role in ethanol CPP comes from knockout studies. Mice deficient in either D1 or D2 receptors as well as mice lacking DARPP-32 (a phosphoprotein regulating D1 receptor function) show reduced ethanol-induced conditioned place preference [81, 82, 83 and 84]. In contrast to these findings, antagonists of the D1 (SCH 23390) or D2 (raclopride, haloperidol) receptor failed to alter the expression of ethanol conditioned place preference (CPP) [85, 86]. The D3-preferring antagonist U99194A was shown to enhance the acquisition of ethanol CPP [87, 88], an

P.143


effect that may depend on an additive interaction with ethanol because this D3 antagonist alone can produce CPP in rats [89]. The facilitating effects of U99194A in pharmacological studies are difficult to reconcile with data from D3 knockout mice that showed no change in ethanol CPP [90]. More importantly, U99194A failed to prevent the expression of ethanol CPP [30] indicating that blockade of D3 receptors does not interfere with the conditioned reinforcing effects of ethanol. In interpreting these CPP data it is important to consider that effects in knockout models may depend on life-long loss of DA receptor function, resulting in behavioral consequences different from those produced by acute DA receptor blockade. Also, use of knockout preparations leaves unclear whether the deficit in the targeted DA receptor population compromises the acquisition of ethanol CPP because ethanol is not reinforcing in these animals, whether these animals are unable to learn Pavlovian associations, or whether deficiency in a particular DA receptor population interferes with the expression of CPP.

As outlined earlier, the CPP data that pertain most directly to the question as to whether DA plays a role in the conditioned incentive effects of ethanol-paired environments are those that have examined whether the expression of CPP is sensitive to pharmacological manipulation of DA neurotransmission. In the studies that have taken this approach, neither D1 nor D2-selective antagonists altered ethanol CPP [85, 86], findings that are in clear contradiction with the attenuation of ethanol-seeking behavior by D1 or D2 antagonists in models of conditioned reinstatement [74], appetitively-motivated responding [78, 79], and conditioned reinforcement [80]. Several factors may account for these discrepancies. First, CPP studies typically employ involuntary administration of ethanol. The reinforcing actions of ethanol under these conditions may differ from those associated with voluntary oral self-administration. As a result, the strength or nature of associations that are formed between ethanol and environmental stimuli may differ in CPP versus self-administration procedures. Importantly, as well, the number of learning trials in models of ethanol-seeking that involve the conditioning of the effects of self-administered ethanol with environmental stimuli is typically greater than in the CPP procedure. Associations that are produced between specific environmental stimuli and ethanol therefore presumably are weaker in the CPP model. Due to these differences, the expression of conditioned ethanol-seeking responses may be differentially sensitive to DA antagonists in CPP vs. self-administration models. Lastly, it cannot be ruled out that different neural substrates are involved in contextual conditioning in the case of CPP as opposed to stimuli associated with active self-administration.

With these considerations in mind, data from behavioral models that involve the conditioning of self-administered ethanol with environmental stimuli consistently confirm that DA antagonists attenuate the motivating effects of ethanol-related environmental stimuli. What limits this information with respect to the potential of DA transmission to serve as a treatment target for the prevention of cue-induced craving and relapse is that the great majority of

P.144


data has been generated in ethanol nondependent animals. In addition to its positive reinforcing effects, alcohol serves as a negative reinforcer during the development of dependence by alleviating aversive withdrawal symptoms (e.g., [91]). This type of drug-related learning may increase the salience of ethanol as a reinforcer and consequently craving or drug-seeking produced by ethanol cues compared to nondependent subjects for which the drug serves as a positive reinforcer only. Additionally, ethanol-seeking in individuals with a history of dependence is likely to be modified by neuroadaptive changes. Chronic ethanol exposure alters DA function and these changes may modify the effects of DA antagonist treatments on ethanol-seeking behavior as illustrated, for example, by the finding that the potency with which D1 and D2 antagonists suppress conditioned reinstatement is increased in previously ethanol-dependent rats [74]. Thus, it will be important to more systematically evaluate the role of DA in conditioned ethanol-seeking using animals with a history of ethanol dependence and withdrawal.

The alcohol deprivation effect

Despite growing evidence of a dopaminergic role in the exacerbation of ethanol intake following deprivation, only few studies have examined whether the ADE is sensitive to pharmacological manipulation of DA receptors. In one study [92], haloperidol administered daily during 14 days of deprivation following six weeks of free access to ethanol reversed the ADE measured on the first post-treatment day in mice. The reduction in drinking was most pronounced during the first 1.5 h of renewed access, a time period during which ethanol intake was greatest in vehicle-treated controls. A proportionally smaller, but significant reduction was observed during the remaining 22.5 h of the first post-deprivation day, suggesting that drinking during the early and late phases of the ADE is differentially sensitive to chronic haloperidol. Nonetheless, the reversal of the ADE during the first hour of renewed access may be indicative of protective effects of chronic D2 antagonist treatments against loss of control. This effect may be mediated via haloperidol-induced upregulation of postsynaptic D2 receptors and consequently enhanced DA transmission, in agreement with the literature implicating D2 deficits in alcohol preference (e.g., [48, 49 and 50]). In a second study, lisuride, an ergot derivative acting as a D2 agonist, failed to reduce ethanol intake in rats tested in a long-term free-access model that utilizes repeated brief deprivation periods to induce high ethanol intake [44]. In fact, lisuride slightly increased consumption in high ethanol drinking as well as previously ethanol-na ve rats, presumably as a result of autoregulatory decreases in DA activity produced by the D2 agonist (see e.g., [93]), a condition that has been linked to increased ethanol preference and intake [48, 49, 50 and 51].

P.145


Table 1.

Model/procedure

Compound/dose

Abstinence

Effect

Species/strain

Reference

Conditioned reinstatement

D1 antagonist SCH 23390 (5, 10, 50 g/kg)

16 days (Extinction)

Decrease

Wistar rats

[54]

D2 antagonist Eticlopride (5, 10, 50 g/kg)

16 days (Extinction)

Decrease

Wistar rats

[54]

Appetitive ethanol-seeking

D2 antagonist Raclopride (1, 3, 10 g/rat; intra-NAcc)

24 h

Decrease

Long-Evans rats

[25]

D2 antagonist Remoxipride (5, 10, 15 mg/kg)

24 h

Decrease

Long-Evans rats

[27]

Conditioned reinforcement

D2 antagonist Haloperidol (0.25 mg/kg)

24 h

Decrease

Sprague-Dawley rats

[113]

Conditioned place preference

D2 antagonist SCH 23390 (0.015, 0.03 mg/kg)

No Effect

DBA/2J mice

[30]

D2 antagonist Raclopride (0.3, 0.6 mg/kg)

No Effect

DBA/2J mice

[30]

D2 antagonist Haloperidol (0.05, 0.1 mg/kg)

No Effect

DBA/2J mice

[22]

D3 antagonist U99194A (10, 20 mg/kg)

No Effect

DBA/2J mice

[30]

D3 antagonist U99194A (10, 20 mg/kg)

Increase

Swiss-Webster mice

[10, 11]

Alcohol deprivation effect

D2 antagonist Haloperidol (1 mg/kg/day for 2 weeks)

14 days

Reduction

CBA C57BL mice

[87]

D2 agonist Lisuride (90 g/kg/day for 8 weeks)

36 weeks

No effect

Wistar rats

[59]

P.146


Is dopamine neurotransmission a promising treatment target?

As elaborated above, the neurobiological literature reveals a significant role for DA in relapse risk associated with alcohol cue exposure as well as neuroadaptive changes in DA function. Preclinical data from studies targeting DA receptors to attenuate ethanol-seeking behavior are largely consistent with this literature and implicate DA neurotransmission as a treatment target. Controlled laboratory studies in humans further support this possibility with findings that blockade of D2 receptors reduces craving induced by an ethanol priming dose, blocks the stimulant and euphorigenic effects of ethanol, and reduces ethanol intake. While the latter findings provide excellent support for a role of DA in acute ethanol reinforcement in humans it remains questionable whether direct pharmacological manipulation of DA transmission represents a promising pharmacotherapeutic approach for relapse prevention.

First, although preclinical evidence strongly suggests a potential for DA antagonists in preventing or ameliorating craving and relapse associated with alcohol cue exposure, the expectation of sustained therapeutic benefits with DA antagonist treatments is fraught with the complication that such treatments would exacerbate the DA hypoactivity that accompanies acute and protracted withdrawal, and that has been linked to increased relapse risk. Second, DA agonist treatments that, based on the preclinical literature, would be predicted to reduce susceptibility to relapse by ameliorating DA hypofunction are problematic as well. The D2 agonist bromocriptine administered during the acute ethanol withdrawal phase reduced craving, anxiety, restlessness and depression in hospitalized alcoholics [41]. Longer bromocriptine treatment (six weeks) reduced craving and anxiety as well, particularly in alcoholics with the D2 receptor A1 allele, but this treatment regimen failed to significantly reduce attrition rates [94]. Similarly, a long-acting bromocriptine preparation (Parlodel-LAR ) proved ineffective in reducing relapse rates in detoxified alcoholics [95], and chronic treatment with another D2 agonist, lisuride, actually enhanced proclivity to relapse by decreasing the latency to resume drinking [93]. This effect, and the lack of anti-relapse efficacy of D2 agonists in general, appears to be a consequence of DA autoregulatory changes associated with chronic DA agonist administration. More specifically, neuroendocrine measures of dopaminergic responsivity to D2 stimulation in alcoholics suggest that chronic lisuride produces an autoinhibitory reduction in dopaminergic tone that likely exacerbates DA hypoactivity, thereby increasing the likelihood of relapse [93].

As more extensively discussed in the Chapter by Carai et al., these clinical findings provide further support for the hypothesis that DA transmission is an important player in the regulation of ethanol-seeking, but illustrate also that direct manipulation of this system, particularly in the context of chronic treatments likely to be required for relapse prevention, is associated with complications that limit therapeutic promise. It remains to be determined whether low-dose D1 antagonist treatments or agents that act as partial agonists at the D2

P.147


receptor may prove beneficial. The latter class of agents that exert functional agonist effects under conditions of low dopaminergic activity but exert antagonist actions when dopaminergic activity is high may offer some promise by ameliorating chronic ethanol-induced DA deficits while at the same time attenuating DA surges associated with ethanol cue exposure. Partial agonists including terguride and SDZ 208-911 reduce ethanol intake in rats with both acute and chronic administration [96], but a possible anti-relapse potential of these agents has not yet been established in appropriate animal models.

To properly evaluate the significance and potential of DA neurotransmission to serve as a target for anti-relapse medications, it is important to bear in mind that indirect modification of DA activity through agents that act on other neurochemical systems may perhaps provide an effective approach, eliminating problematic side- and secondary effects associated with chronic DA agonist or antagonist treatments. In fact, as briefly discussed below, regulation or modification of DA activity by other neurochemical systems is likely to contribute to the therapeutic actions of agents with established anti-craving or anti-relapse efficacy.

Effects of indirect manipulation of dopamine transmission on ethanol-seeking

Opiate antagonists

The nonselective opiate receptor antagonist naltrexone as well as mu-opioid and delta-opioid-selective antagonists effectively reverse the behavioral effects of ethanol-related stimuli in conditioned reinstatement [68, 70, 71, 97] and CPP [98, 99] tests. Paralleling these findings, naltrexone attenuates cue-induced ethanol craving [100, 101] and lowers relapse rates in alcoholic populations ([102, 103, 104, 105, 106 and 107]; see also Chapters by Cowen and O'Brien et al.).

Opiate antagonists blunt the stimulatory effects of systemically and self-administered ethanol on DA release in the nucleus accumbens [10, 108, 109 and 110]. Interference with ethanol-induced activation of mesolimbic DA transmission and, consequently, ethanol reward has therefore been implicated as a mechanism by which opiate antagonists reduce ethanol intake. This naltrexone-induced decrease in the rewarding efficacy of ethanol also may limit ethanol intake during a lapse in abstinent alcoholics and thereby lower the incidence of full-blown relapse.

A second mechanism for the therapeutic effects of naltrexone involving interactions with DA transmission may be interference with the neurochemical effects of alcohol cues hypothesized to underlie conditioned cue reactivity and craving. Anticipation of ethanol and exposure to alcohol-related contextual stimuli can increase extracellular DA levels in the nucleus accumbens [5, 10, 11 and 12], and alcohol cue exposure activates DA-rich brain regions in abstinent alcoholics [27]. In view of the ability of opiate antagonists to block ethanol-induced

P.148


stimulation of DA release, it is possible that these agents also interfere with the DA-activating actions of ethanol cues. Direct evidence supporting this hypothesis is still lacking. However, it is not unlikely that such effects may contribute to the anti-craving effects of opiate antagonists in humans and their inhibitory effects on conditioned ethanol-seeking behavior in animals.

Acamprosate

Acamprosate reduces relapse during ethanol withdrawal in humans and lowers ethanol-seeking and intake in animal models (e.g., [111, 112]; see also Chapters by de Witte et al. and Mann). The mechanism(s) underlying these effects are not well understood, but are thought to involve reduction of neural hyperexcitability associated with ethanol withdrawal via restoration of normal N-methyl-D-aspartate receptor function [112, 113 and 114]. Recent evidence indicates, however, that acute acamprosate treatment also dose-dependently delays and suppresses ethanol's stimulatory effects on DA release in the nucleus accumbens [115]. This suggests that acamprosate may reduce relapse during withdrawal by attenuating the ability of ethanol to activate the mesolimbic DA reward pathway. Acamprosate, however, did not attenuate appetitive-phase responding (i.e., ethanol-seeking elicited and maintained by the incentive-motivational effects of a drug-related environment) in rats [116]. Thus, the DA inhibitory effect of acamprosate and its consequences on ethanol intake may be limited to the direct pharmacological actions of ethanol, without interfering with the presumably DA-mediated effects of alcohol-related environmental stimuli on ethanol-seeking behavior and craving.

Acknowledgements

The author would like to thank Mike Arends for his editorial assistance.

References

1 Brodie MS, Shefner SA, Dunwiddie TV (1990) Ethanol increases the firing rate of dopamine neurons of the rat ventral tegmental area in vitro. Brain Res 508: 65-69

2 Brodie MS, Pesold C, Appel SB (1999) Ethanol directly excites dopaminergic ventral tegmental area reward neurons. Alcohol Clin Exp Res 23: 1848-1852

3 Gessa GL, Muntoni F, Collu M, Vargiu L, Mereu G (1985) Low doses of ethanol activate dopaminergic neurons in the ventral tegmental area. Brain Res 348: 201-203

4 Imperato A, DiChiara G (1986) Preferential stimulation of dopamine release in the nucleus accumbens of freely moving rats by ethanol. J Pharmacol Exp Ther 239: 219-239

5 Weiss F, Lorang MT, Bloom FE, Koob GF (1993) Oral alcohol self-administration stimulates dopamine release in the rat nucleus accumbens: Genetic and motivational determinants. J Pharmacol Exp Ther 267: 250-258

6 Koob GF, Vaccarino F, Amalric M, Bloom FE (1987) Positive reinforcement properties of drugs: Search for neural substrates. In: J Engel, L Oreland, DH Ingvar, B Pernow, S Rossner, LA Pellborn (eds): Brain Reward Systems and Abuse. Raven Press, New York, 35-50

7 Samson HH (1992) The function of brain dopamine in ethanol reinforcement. In: RR Watson (ed.): Alcohol and Neurobiology: Receptors, Membranes, and Channels. CRC Press, Boca Raton, FL, 91-107

P.149


8 Spanagel R, Weiss F (1999) The dopamine hypothesis of reward: past and current status. Trends Neurosci 22: 521-527

9 Wise RA, Rompre PP (1989) Brain dopamine and reward. Annu Rev Psychol 40: 191-225

10 Gonzales RA, Weiss F (1998) Suppression of ethanol-reinforced behavior by naltrexone is associated with attenuation of the ethanol-induced increase in dialysate dopamine levels in the nucleus accumbens. J Neurosci 18: 10663-10671

11 Melendez RI, Rodd-Henricks ZA, Engleman EA, Li TK, McBride WJ, Murphy JM (2002) Microdialysis of dopamine in the nucleus accumbens of alcohol-preferring (P) rats during anticipation and operant self-administration of ethanol. Alcohol Clin Exp Res 26: 318-325

12 Vavrousek-Jakuba E, Cohen CA, Shoemaker WJ (1992) Ethanol effects on dopamine receptors: In vivo binding following voluntary ethanol (ETOH) intake in rats. In: CA Naranjo, EM Sellers (eds: Novel Pharmacological Interventions for Alcoholism. Springer, New York, 372-374

13 O'Brien CP, Childress AR, Ehrman R, Robbins SJ (1998) Conditioning factors in drug abuse: Can they explain compulsion? J Psychopharmacol 12: 15-22

14 Cooney NL, Litt MD, Morse PA, Bauer LO, Gaupp L (1997) Alcohol cue reactivity, negativemood reactivity, and relapse in treated alcoholic men. J Abnorm Psychol 106: 243-250

15 Kaplan RF, Cooney NL, Baker LH, Gillespie RA, Meyer RE, Pomerleau OF (1985) Reactivity to alcohol-related cues: Physiological and subjective responses in alcoholics and nonproblem drinkers. J Stud Alcohol 46: 267-272

16 Laberg JC (1986) Alcohol and expectancy: Subjective, psychophysiological and behavioral responses to alcohol stimuli in severely, moderately and non-dependent drinkers. Br J Addict 81: 797-808

17 McCusker CG, Brown K (1990) Alcohol-predictive cues enhance tolerance to and precipitate craving for alcohol in social drinkers. J Stud Alcohol 51: 494-499

18 McCusker CG, Brown K (1995) Cue-exposure to alcohol-associated stimuli reduces autonomic reactivity, but not craving and anxiety, in dependent drinkers. Alcohol Alcoholism 30: 319-327

19 Monti PM, Binkoff JA, Abrams DB, Zwick WR, Nirenberg TD, Liepman MR (1987) Reactivity of alcoholics and nonalcoholics to drinking cues. J Abnorm Psychol 96: 122-126

20 Monti PM, Rohsenow DJ, Rubonis AV, Niaura RS, Sirota AD, Colby SM, Abrams DB (1993) Alcohol cue reactivity: effects of detoxification and extended exposure. J Stud Alcohol 54: 235-245

21 Staiger PK, White JM (1991) Cue reactivity in alcohol abusers: Stimulus specificity and extinction of the responses. Addict Behav 16: 211-221

22 Miller NS, Gold MS (1994) Dissociation of conscious desire (craving) from and relapse in alcohol and cocaine dependence. Ann Clin Psychiat 6: 99-106

23 Tiffany ST, Carter BL (1998) Is craving the source of compulsive drug use? J Psychopharmacol 12: 23-30

24 Everitt BJ, Wolf ME (2002) Psychomotor stimulant addiction: a neural systems perspective. J Neurosci 22: 3312-3320

25 Ikemoto S, Panksepp J (1999) The role of nucleus accumbens dopamine in motivated behavior: a unifying interpretation with special reference to reward-seeking. Brain Res Brain Res Rev 31: 6-41

26 Robbins TW, Everitt BJ (1996) Neurobehavioural mechanisms of reward and motivation. Curr Opin Neurobiol 6: 228-236

27 Braus DF, Wrase J, Grusser S, Hermann D, Ruf M, Flor H, Mann K, Heinz A (2001) Alcohol-associated stimuli activate the ventral striatum in abstinent alcoholics. J Neural Transm 108: 887-894

28 Slawecki CJ, Hodge CW, Samson HH (1997) Dopaminergic and opiate agonists and antagonists differentially decrease multiple schedule responding maintained by sucrose/ethanol and sucrose. Alcohol 14: 281-294

29 Jaffe JH, Cascella NG, Kumor KM, Sherer MA (1989) Cocaine-induced cocaine craving. Psychopharmacology 97: 59-64

30 Ludwig AM, Wikler A, Stark LH (1974) The first drink: psychobiological aspects of craving. Arch Gen Psychiatry 30: 539-547

31 Modell JG, Mountz JM, Glaser FB, Lee JY (1993) Effect of haloperidol on measures of craving and impaired control in alcoholic subjects. Alcohol Clin Exp Res 17: 234-240

32 Enggasser JL, de Wit H (2001) Haloperidol reduces stimulant and reinforcing effects of ethanol in social drinkers. Alcohol Clin Exp Res 25: 1448-1456

P.150


33 Fadda F, Argiolas A, Melis MR, Serra G, Gessa GL (1980) Differential effect of acute and chronic ethanol on dopamine metabolism in frontal cortex, caudate nucleus and substantia nigra. Life Sci 27: 979-986

34 Rothblat DS, Rubin E, Schneider JS (2001) Effects of chronic alcohol ingestion on the mesostriatal dopamine system in the rat. Neurosci Lett 300: 63-66

35 Tabakoff B, Hoffman PL (1978) Alterations in receptors controlling dopamine synthesis after chronic ethanol ingestion. J Neurochem 31: 1223-1229

36 Diana M, Pistis M, Carboni S, Gessa GL, Rossetti ZL (1993) Profound decrement of mesolimbic neuronal activity during ethanol withdrawal syndrome in rats: Electrophysiological and biochemical evidence. Proc Natl Acad Sci USA 90: 7966-7969

37 Shen RY, Chiodo LA (1993) Acute withdrawal after repeated ethanol treatment reduces the number of spontaneously active dopaminergic neurons in the ventral tegmental area. Brain Res 622: 289-293

38 Rossetti ZL, Hmaidan Y, Gessa GL (1992) Marked inhibition of mesolimbic dopamine release: A common feature of ethanol, morphine, cocaine and amphetamine abstinence in rats. Eur J Pharmacol 221: 227-234

39 Weiss F, Parsons LH, Schulteis G, Hyyti P, Lorang MT, Bloom FE, Koob GF (1996) Ethanol self-administration restores withdrawal-associated deficiencies in accumbal dopamine and 5-hydroxytryptamine release in dependent rats. J Neurosci 16: 3474-3485

40 Rossetti ZL, Melis F, Carboni S, Diana M, Gessa GL (1992) Alcohol withdrawal in rats is associated with a marked fall in extraneuronal dopamine. Alcohol Clin Exp Res 16: 529-532

41 Borg V, Weinholdt T (1982) Bromocriptine in the treatment of the alcohol-withdrawal syndrome. Acta Psychiatr Scand 65: 101-111

42 Diana M, Pistis M, Muntoni A, Gessa G (1996) Mesolimbic dopaminergic reduction outlasts ethanol withdrawal syndrome: evidence of protracted abstinence. Neuroscience 71: 411-415

43 Bailey CP, Andrews N, McKnight AT, Hughes J, Little HJ (2000) Prolonged changes in neurochemistry of dopamine neurones after chronic ethanol consumption. Pharmacol Biochem Behav 66: 153-161

44 May T, Wolf U, Wolffgramm J (1995) Striatal dopamine receptors and adenylyl cyclase activity in a rat model of alcohol addiction: effects of ethanol and lisuride treatment. J Pharmacol Exp Ther 275: 1195-1203

45 Heinz A, Dettling M, Kuhn S, Dufeu P, Graf KJ, Kurten I, Rommelspacher H, Schmidt IG (1995) Blunted growth hormone response is associated with early relapse in alcohol-dependent patients. Alcohol Clin Exp Res 19: 62-65

46 Heinz A, Lichtenberg-Kraag B, Baum SS, Graf K, Kruger F, Dettling M, Rommelspacher H (1995) Evidence for prolonged recovery of dopaminergic transmission after detoxification in alcoholics with poor treatment outcome. J Neural Transm Gen Sect 102: 149-157

47 Wiesbeck GA, Weijers HG, Gross JP (2000) Craving for alcohol and dopamine receptor sensitivity in alcohol-dependent men and control subjects. J Neural Transm 107: 691-699

48 George SR, Fan T, Ng GY, Jung SY, O'Dowd BF, Naranjo CA (1995) Low endogenous dopamine function in brain predisposes to high alcohol preference and consumption: Reversal by increasing synaptic dopamine. J Pharmacol Exp Ther 273: 373-379

49 McBride WJ, Murphy JM, Gatto GJ, Levy AD, Lumeng L, Li TK (1991) Serotonin and dopamine systems regulating alcohol intake. Alcohol Alcoholism Suppl 1: 411-416

50 McBride WJ, Li TK (1998) Animal models of alcoholism: neurobiology of high alcohol-drinking behavior in rodents. Crit Rev Neurobiol 12: 339-369

51 Thanos PK, Volkow ND, Freimuth P, Umegaki H, Ikari H, Roth G, Ingram DK, Hitzemann R (2001) Overexpression of dopamine D2 receptors reduces alcohol self-administration. J Neurochem 78: 1094-1103

52 Sinclair JD, Senter RJ (1967) Increased preference for ethanol in rats following alcohol deprivation. Psychonomic Sci 8: 11-12

53 Sinclair JD, Li TK (1989) Long and short alcohol deprivation: Effects on AA and P alcohol-preferring rats. Alcohol 6: 505-509

54 Rankin HR, Hodgson R, Stockwell T (1979) The concept of craving and its measurement. Behav Res Ther 17: 389-396

55 Wolffgramm J, Heyne A (1995) From controlled drug intake to loss of control: the irreversible development of drug addiction in the rat. Behav Brain Res 70: 77-94

P.151


56 Kornet M, Goosen C, Van Ree JM (1991) Effects of naltrexone on alcohol consumption during chronic alcohol drinking and after a period of imposed abstinence in free-choice drinking rhesus monkeys. Psychopharmacology 104: 367-376

57 Burish TG, Maisto SA, Cooper AM, Sobell MB (1981) Effects of voluntary short-term abstinence from alcohol on subsequent drinking patterns of college students. J Stud Alcohol 42: 1013-1020

58 Ludwig AM, Wikler A (1974) Craving and relapse to drink. Q J Stud Alcohol 35: 108-130

59 O'Donnell PJ (1984) The abstinence violation effect and circumstances surrounding relapse as predictors of outcome status in male alcoholic outpatients. J Psychol 117: 257-262

60 Spanagel R, H lter SM, Allingham K, Landgraf R, Zieglg nsberger W (1996) Acamprosate and alcohol: I. Effects on alcohol intake following alcohol deprivation in the rat. Eur J Pharmacol 305: 39-44

61 Wolffgramm J (1991) An ethopharmacological approach to the development of drug addiction. Neurosci Biobehav Rev 15: 515-519

62 Thielen RJ, Engleman EA, Rodd ZA, Murphy JM, Lumeng L, Li TK, McBride WJ (2004) Ethanol drinking and deprivation alter dopaminergic and serotonergic function in the nucleus accumbens of alcohol-preferring rats. J Pharmacol Exp Ther 309: 216-225

63 Nestby P, Vanderschuren LJ, De Vries TJ, Mulder AH, Wardeh G, Hogenboom F, Schoffelmeer AN (1999) Unrestricted free-choice ethanol self-administration in rats causes long-term neuroadaptations in the nucleus accumbens and caudate putamen. Psychopharmacology (Berl) 141: 307-314

64 L A, Shaham Y (2002) Neurobiology of relapse to alcohol in rats. Pharmacol Ther 94: 137-156

65 Shaham Y, Shalev U, Lu L, De Wit H, Stewart J (2003) The reinstatement model of drug relapse: history, methodology and major findings. Psychopharmacology (Berl) 168: 3-20

66 Bienkowski P, Kostowski W, Koros E (1999) Ethanol-reinforced behaviour in the rat: effects of naltrexone. Eur J Pharmacol 374: 321-327

67 Bienkowski P, Kostowski W, Koros E (1999) The role of drug-paired stimuli in extinction and reinstatement of ethanol-seeking behaviour in the rat. Eur J Pharmacol 374: 315-319

68 Ciccocioppo R, Martin-Fardon R, Weiss F (2002) Effect of selective blockade of mu(1) or delta opioid receptors on reinstatement of alcohol-seeking behavior by drug-associated stimuli in rats. Neuropsychopharmacology 27: 391-399

69 Ciccocioppo R, Economidou D, Fedeli A, Angeletti S, Weiss F, Heilig M, Massi M (2003) Attenuation of ethanol self-administration and of conditioned reinstatement of alcohol-seeking behaviour by the antiopioid peptide nociceptin/orphanin FQ in alcohol-preferring rats. Psychopharmacology (Berl) 172: 170-178

70 Ciccocioppo R, Lin D, Martin-Fardon R, Weiss F (2003) Reinstatement of ethanol-seeking behavior by drug cues following single versus multiple ethanol intoxication in the rat: effects of naltrexone. Psychopharmacology (Berl) 168: 208-215

71 Katner SN, Magalong JG, Weiss F (1999) Reinstatement of alcohol-seeking behavior by drug-associated discriminative stimuli after prolonged extinction in the rat. Neuropsychopharmacology 20: 471-479

72 Katner SN, Weiss F (1999) Ethanol-associated olfactory stimuli reinstate ethanol-seeking behavior after extinction and modify extracellular dopamine levels in the nucleus accumbens. Alcohol Clin Exp Res 23: 1751-1760

73 Ciccocioppo R, Angeletti S, Weiss F (2001) Long-lasting resistance to extinction of response reinstatement induced by ethanol-related stimuli: role of genetic ethanol preference. Alcohol Clin Exp Res 25: 1414-1419

74 Liu X, Weiss F (2002) Reversal of ethanol-seeking behavior by D1 and D2 antagonists in an animal model of relapse: differences in antagonist potency in previously ethanol-dependent versus nondependent rats. J Pharmacol Exp Ther 300: 882-889

75 Samson HH, Slawecki CJ, Sharpe AL, Chappell A (1998) Appetitive and consummatory behaviors in the control of ethanol consumption: a measure of ethanol seeking behavior. Alcohol Clin Exp Res 22: 1783-1787

76 Samson HH, Sharpe AL, Denning C (1999) Initiation of ethanol self-administration in the rat using sucrose substitution in a sipper-tube procedure. Psychopharmacology (Berl) 147: 274-279

77 Samson HH, Czachowski CL, Slawecki CJ (2000) A new assessment of the ability of oral ethanol to function as a reinforcing stimulus. Alcohol Clin Exp Res 24: 766-773

78 Czachowski CL, Chappell AM, Samson HH (2001) Effects of raclopride in the nucleus accumbens on ethanol seeking and consumption. Alcohol Clin Exp Res 25: 1431-1440

79 Czachowski CL, Santini LA, Legg BH, Samson HH (2002) Separate measures of ethanol seeking and drinking in the rat: effects of remoxipride. Alcohol 28: 39-46

P.152


80 Wilson AW, Costall B, Neill JC (2000) Manipulation of operant responding for an ethanol-paired conditioned stimulus in the rat by pharmacological alteration of the serotonergic system. J Psychopharmacol 14: 340-346

81 Cunningham CL, Howard MA, Gill SJ, Rubinstein M, Low MJ, Grandy DK (2000) Ethanol-conditioned place preference is reduced in dopamine D2 receptor-deficient mice. Pharmacol Biochem Behav 67: 693-699

82 El-Ghundi M, George SR, Drago J, Fletcher PJ, Fan T, Nguyen T, Liu C, Sibley DR, Westphal H, O'Dowd BF (1998) Disruption of dopamine D1 receptor gene expression attenuates alcohol-seeking behavior. Eur J Pharmacol 353: 149-158

83 Risinger FO, Freeman PA, Rubinstein M, Low MJ, Grandy DK (2000) Lack of operant ethanol self-administration in dopamine D2 receptor knockout mice. Psychopharmacology (Berl) 152: 343-350

84 Risinger FO, Freeman PA, Greengard P, Fienberg AA (2001) Motivational effects of ethanol in DARPP-32 knock-out mice. J Neurosci 21: 340-348

85 Cunningham CL, Malott DH, Dickinson SD, Risinger FO (1992) Haloperidol does not alter expression of ethanol-induced conditioned place preference. Behav Brain Res 50: 1-5

86 Dickinson SD, Lee EL, Rindal K, Cunningham CL (2003) Lack of effect of dopamine receptor blockade on expression of ethanol-induced conditioned place preference in mice. Psychopharmacology (Berl) 165: 238-244

87 Boyce JM, Risinger FO (2000) Enhancement of ethanol reward by dopamine D3 receptor blockade. Brain Res 880: 202-206

88 Boyce JM, Risinger FO (2002) Dopamine D3 receptor antagonist effects on the motivational effects of ethanol. Alcohol 28: 47-55

89 Kling-Petersen T, Ljung E, Wollter L, Svensson K (1995) Effects of dopamine D3 preferring compounds on conditioned place preference and intracranial self-stimulation in the rat. J Neural Transm Gen Sect 101: 27-39

90 Boyce-Rustay JM, Risinger FO (2003) Dopamine D3 receptor knockout mice and the motivational effects of ethanol. Pharmacol Biochem Behav 75: 373-379

91 Roberts AJ, Heyser CJ, Cole M, Griffin P, Koob GF (2000) Excessive ethanol drinking following a history of dependence: animal model of allostasis. Neuropsychopharmacology 22: 581-594

92 Salimov RM, Salimova NB, Shvets LN, Maisky AI (2000) Haloperidol administered subchronically reduces the alcohol-deprivation effect in mice. Alcohol 20: 61-68

93 Schmidt LG, Kuhn S, Smolka M, Schmidt K, Rommelspacher H (2002) Lisuride, a dopamine D2 receptor agonist, and anticraving drug expectancy as modifiers of relapse in alcohol dependence. Prog Neuropsychopharmacol Biol Psychiatry 26: 209-217

94 Lawford BR, Young RM, Rowell JA, Qualichefski J, Fletcher BH, Syndulko K, Ritchie T, Noble EP (1995) Bromocriptine in the treatment of alcoholics with the D2 dopamine receptor A1 allele. Nat Med 1: 337-341

95 Naranjo CA, Dongier M, Bremner KE (1997) Long-acting injectable bromocriptine does not reduce relapse in alcoholics. Addiction 92: 969-978

96 Bono G, Balducci C, Richelmi P, Koob GF, Pulvirenti L (1996) Dopamine partial receptor agonists reduce ethanol intake in the rat. Eur J Pharmacol 296: 233-238

97 Liu X, Weiss F (2002) Additive effect of stress and drug cues on reinstatement of ethanol seeking: exacerbation by history of dependence and role of concurrent activation of corticotropin-releasing factor and opioid mechanisms. J Neurosci 22: 7856-7861

98 Cunningham CL, Henderson CM, Bormann NM (1998) Extinction of ethanol-induced conditioned place preference and conditioned place aversion: effects of naloxone. Psychopharmacology (Berl) 139: 62-70

99 Gerrits MAFM, Patkina N, Zvartau EE, van Ree JM (1995) Opioid blockade attenuates acquisition and expression of cocaine-induced place preference conditioning in rats. Psychopharmacology (Berl) 119: 92-98

100 Monti PM, Rohsenow DJ, Hutchison KE, Swift RM, Mueller TI, Colby SM, Brown RA, Gulliver SB, Gordon A, Abrams DB (1999) Naltrexone's effect on cue-elicited craving among alcoholics in treatment. Alcohol Clin Exp Res 23: 1386-1394

101 Rohsenow DJ, Monti PM, Hutchison KE, Swift RM, Colby SM, Kaplan GB (2000) Naltrexone's effects on reactivity to alcohol cues among alcoholic men. J Abnorm Psychol 109: 738-742

P.153


102 Anton RF, Moak DH, Waid LR, Latham PK, Malcolm RJ, Dias JK (1999) Naltrexone and cognitive behavioral therapy for the treatment of outpatient alcoholics: results of a placebo-controlled trial. Am J Psychiatry 156: 1758-1764

103 Anton RF, Moak DH, Latham PK, Waid LR, Malcolm RJ, Dias JK, Roberts JS (2001) Posttreatment results of combining naltrexone with cognitive-behavior therapy for the treatment of alcoholism. J Clin Psychopharmacol 21: 72-77

104 Monterosso JR, Flannery BA, Pettinati HM, Oslin DW, Rukstalis M, O'Brien CP, Volpicelli JR (2001) Predicting treatment response to naltrexone: the influence of craving and family history. Am J Addict 10: 258-268

105 O'Brien CP, Volpicelli LA, Volpicelli JR (1996) Naltrexone in the treatment of alcoholism: a clinical review. Alcohol 13: 35-39

106 O'Malley SS, Jaffe AJ, Chang G, Schottenfeld RS, Meyer RE, Rounsaville B (1992) Naltrexone and coping skills therapy for alcohol dependence: a controlled study. Arch Gen Psychiatry 49: 881-887

107 Volpicelli JR, Alterman AI, Hayashida M, O'Brien CP (1992) Naltrexone in the treatment of alcohol dependence. Arch Gen Psychiatry 49: 876-880

108 Acquas E, Meloni M, Di Chiara G (1993) Blockade of -opioid receptors in the nucleus accumbens prevents ethanol-induced stimulation of dopamine release. Eur J Pharmacol 230: 239-241

109 Benjamin D, Grant ER, Pohorecky LA (1993) Naltrexone reverses ethanol-induced dopamine release in the nucleus accumbens in awake, freely moving rats. Brain Res 621: 137-140

110 Widdowson PS, Holman RB (1992) Ethanol-induced increase in endogenous dopamine release may involve endogenous opiates. J Neurochem 59: 157-163

111 Mason BJ, Ownby RL (2000) Acamprosate for the treatment of alcohol dependence: a review of double-blind, placebo-controlled trials. CNS Spectrums 5: 58-69

112 Spanagel R, Zieglgansberger W (1997) Anti-craving compounds for ethanol: new pharmacological tools to study addictive processes. Trends Pharmacol Sci 18: 54-59

113 Littleton J (1995) Acamprosate in alcohol dependence: How does it work? Addiction 90: 1179-1188

114 Wilde MI, Wagstaff AJ (1997) Acamprosate. A review of its pharmacology and clinical potential in the management of alcohol dependence after detoxification. Drugs 53: 1038-1053

115 Olive MF, Nannini MA, Ou CJ, Koenig HN, Hodge CW (2002) Effects of acute acamprosate and homotaurine on ethanol intake and ethanol-stimulated mesolimbic dopamine release. Eur J Pharmacol 437: 55-61

116 Czachowski CL, Legg BH, Samson HH (2001) Effects of acamprosate on ethanol-seeking and self-administration in the rat. Alcohol Clin Exp Res 25: 344-350



Drugs for Relapse Prevention of Alcoholism
Drugs for Relapse Prevention of Alcoholism (Milestones in Drug Therapy)
ISBN: 3764302143
EAN: 2147483647
Year: 2005
Pages: 26

flylib.com © 2008-2017.
If you may any questions please contact us: flylib@qtcs.net