Skip to main content
British Journal of Clinical Pharmacology logoLink to British Journal of Clinical Pharmacology
. 2014 Jan 22;77(2):315–323. doi: 10.1111/bcp.12070

The clinical pharmacology of acamprosate

Nicola J Kalk 1,, Anne R Lingford-Hughes 1
PMCID: PMC4014018  PMID: 23278595

Abstract

Acamprosate is one of the few medications licensed for prevention of relapse in alcohol dependence, and over time it has proved to be significantly, if moderately, effective, safe and tolerable. Its use is now being extended into other addictions and neurodevelopmental disorders. The mechanism of action of acamprosate has been less clear, but in the decade or more that has elapsed since its licensing, a body of translational evidence has accumulated, in which preclinical findings are replicated in clinical populations. Acamprosate modulates N-methyl-d-aspartic acid receptor transmission and may have indirect effects on γ-aminobutyric acid type A receptor transmission. It is known to decrease brain glutamate and increase β-endorphins in rodents and man. Acamprosate diminishes reinstatement in ethanolized rodents and promotes abstinence in humans. Although acamprosate has been called an anticraving drug, its subjective effects are subtle and relate to diminished arousal, anxiety and insomnia, which parallel preclinical findings of decreased withdrawal symptoms in animals treated with acamprosate. Further understanding of the pharmacology of acamprosate will allow appropriate targeting of therapy in individuals with alcohol dependence and extension of its use to other addictions.

Keywords: acamprosate, alcohol dependence, pharmacology

Introduction

Acamprosate can provide effective pharmacotherapy for prevention of relapse in alcohol dependence. Although this has been known for almost two decades and has been confirmed in several meta-analyses, evidence regarding its mechanism of action has accrued more slowly. In 1995, John Littleton proposed that acamprosate prevents relapse in alcohol dependence by decreasing subclinical withdrawal symptoms that may cause an individual to crave alcohol for relief, i.e. ‘negative reinforcement’ [1]. Its effectiveness in reducing measurable alcohol craving has been less impressive. Here, we summarize the evidence and consider to what extent it is consistent with acamprosate suppressing craving related to negative reinforcement.

Acamprosate pharmacokinetics

Acamprosate is formulated as a dimer of acetyl-homotaurine linked by a calcium salt, adjustments made to facilitate absorption. Nonetheless, oral bioavailability remains low, at around 11% [2]. In fact, the absorption of acamprosate is so slow that it acts as though it were a modified-release medication [3], with a time at the maximal concentration of 6.3 h in enteric-coated form [2]. It has a half-life of around 32 h [2], with complete elimination occurring only 4 days after cessation of therapy [4]. The steady-state plasma concentration, reached by around day 5 of treatment, varies from around 1.5 to 5 μm according to research study [2,46]. Acamprosate pharmacokinetics is not affected by alcohol, benzodiazepines or disulfiram 2, but its plasma levels are increased by naltrexone [2,5] and decreased by co-administration with food [2].

Acamprosate is not protein bound, and it is completely dissociated in plasma. It is not metabolized and is excreted unchanged in the urine. Predictably, renal impairment is related to increased plasma concentrations of acamprosate [2]; for this reason, acamprosate is contraindicated in severe renal impairment and dose reduction is advised in mild renal impairment. Plasma levels in Child's A and B liver disease are within normal limits [2]. Furthermore, there has been no increase in markers of liver dysfunction in clinical trials of acamprosate, and no increase in adverse events was found in those participants with more severely deranged liver function tests at baseline [7]. Indeed, liver function tests usually improve with decreased drinking associated with acamprosate therapy.

Receptor pharmacology

There have been few studies of the receptor binding properties of acamprosate. Those that exist concern the N-methyl-d-aspartic acid (NMDA) receptor. Acamprosate has a very low affinity for the dizocilpine and spermidine sites on the NMDA receptor, >1 mm and 600 μm, respectively [8,9]. It has been suggested that any inhibition of binding occurs via allosteric interaction rather than direct competition [9,10].

Mechanism of action

Effect on receptors

More information about the action of acamprosate at receptors has been elucidated through electrophysiological studies. Acamprosate was initially thought to modulate γ-aminobutyric acid type A (GABAA) receptors [11], but in vitro evidence is consistent with an indirect effect. Acamprosate had no effect on GABAA-mediated inhibitory postsynaptic currents in the rat nuclear accumbens (300 μm) [12], on rat cortical neurons (up to 1 mm) [13] or on recombinant human GABAA receptors (up to 100 μm) [14]. However, decreases in burst amplitude and frequency induced by acamprosate (100–400 μm) were partly reversed by bicuculline [15], and increased GABA binding in the hippocampus and thalamus has been observed after administration of acamprosate [16]. Acamprosate may influence GABAA transmission via inhibition of presynaptic γ-aminobutyric acid type B (GABAB) receptors [12].

Likewise, the effect of acamprosate on NMDA receptors may be indirect. Although early electrophysiology studies showed an increase in NMDA excitation in the nucleus accumbens and hippocampus following acamprosate treatment [12,17], this is likely to have resulted from the calcium rather than acamprosate itself [18]. Antagonism of NMDA receptors has been reported subsequently in the cortex, hippocampus and midbrain [1921]. The effect of acamprosate on NMDA receptors may be mediated by interaction at the polyamine site. It reversed spermidine-related potentiation in a subset of striatal neurons only [19] and prevented polyamine-related neurotoxicity in hippocampal slice culture at a clinically relevant concentration (200 nm) [18]. Its effect on NMDA receptors may also be influenced by alcohol, because al Qatari and colleagues reported a biphasic influence on dizopilcine binding to the NMDA pore, with potentiation at low doses and inhibition at high doses in naïve rats but only inhibition in rats subjected to chronic inhalation of ethanol [8].

Acamprosate may also modulate the NMDA receptor response via metabotropic glutamate receptor subtype 5 (mGluR5) antagonism. Acamprosate competes with 1-aminocyclopentane-trans-1,3-dicarboxylic acid (trans-ACPD), a type I metabotropic glutamate receptor (mGluR) antagonist, and prevents trans-ACPD-related neurotoxicity [10]. The attenuation of the effects and withdrawal symptoms of alcohol by acamprosate was similar to that of methylphenylethynylpyridine, a known mGluR5 antagonist [22]. It does not attenuate alcohol intoxication or withdrawal in mGluR5 knock-out mice, suggesting that this receptor is essential for these actions [22]. Likewise, acamprosate may act to increase dopamine in the nucleus accumbens via an indirect pathway involving both mGluR5 and glycine receptors [23]. Preliminary results indicate that acamprosate may be of therapeutic value in fragile X syndrome, which is characterized by type 1 mGluR overactivity [24].

Effects on neurotransmitter concentrations

Acamprosate (400 mg kg−1) prevents the increase in glutamate in the nucleus accumbens of chronically alcoholized rats during alcohol withdrawal [25] and the escalation in the hyperglutamatergic state that occurs with repeated withdrawal episodes [26]. Acamprosate has been associated with an increase in β-endorphin in rats, which is more marked in alcohol-preferring rats [27].

Clinical studies: effect on neurotransmitters and hormones

Preclinical effects on neurotransmitters have been replicated in man. In recently abstinent human alcohol-dependent patients, acamprosate treatment was associated with a decrease in frontal lobe glutamate as measured by magnetic resonance spectroscopy, whereas an increase was seen in a comparable placebo group [28]. Likewise, one study reported that acamprosate was associated with higher plasma β-endorphin in those with higher than median alcohol consumption prior to detoxification [29]. However, the increase in β-endorphin following acamprosate therapy has not been replicated in an independent sample, which may reflect the shorter treatment time or that the sample was not separated into different groups according to consumption [30].

The effect of acamprosate on the hypothalamo-pituitary–adrenal axis is uncertain. Although a trial of acamprosate therapy resulted in persistently raised plasma adrenocorticotrophic hormone (ACTH) and cortisol levels, whereas these normalized in the placebo group [29], other studies do not support an effect of acamprosate on the hypothalamo-pituitary–adrenal axis. A single intravenous dose had no effect on pituitary hormone secretion [31]. Treatment with acamprosate for 3 weeks resulted in no change in plasma cortisol relative to placebo [30]. There was no difference in ACTH or cortisol response to a corticotrophin-releasing hormone stimulation test in alcohol-dependent patients after 1 week of acamprosate with respect to placebo [32]. It may be that the effect is delayed, because the single positive study measured differences in hormones at 4, 8 and 12 weeks.

Relevance of the pharmacology of acamprosate to alcohol dependence

In summary, acamprosate has been found to modulate NMDA receptor transmission and GABAA transmission, to decrease glutamate during alcohol withdrawal, to increase β-endorphin in those with very high alcohol intake and, perhaps, to modulate the hypothalamo-pituitary–adrenal axis. The relevance of the actions of acamprosate on glutamate, NMDA and GABAA transmission to alcohol dependence is that they would compensate for the neurobiological derangement produced by alcohol withdrawal. Acute alcohol potentiates the effect of GABA on GABAA receptors and decreases NMDA, AMPA and kainate receptor function (for a review, see [33]). With chronic alcohol intake, GABAA receptors are downregulated and rendered less responsive by expression of different subtypes 34, whereas NMDA receptors are upregulated [35]. These compensatory changes produce a hyperglutamatergic, potentially excitotoxic state during alcohol withdrawal [26]. The pharmacological profile of acamprosate is therefore consistent with a medication that could target subclinical withdrawal symptoms and, indeed, even offer neuroprotection during withdrawal.

Behavioural effects in animals

Consistent with its pharmacological actions, acamprosate decreases signs of alcohol withdrawal, such as hypermobility [3638] and anxiety-like behaviour [39,40], at doses from 50 mg kg−1 day−1 in chronically ethanolized rodents. Importantly, acamprosate reduces more severe withdrawal complications, such as handling-induced seizures in mice, with a similar reduction at a 100 mg kg−1 dose to that seen with a diazepam dose of 0.25 mg kg−1 [41]. Some physical consequences of withdrawal, for example, hypothermia, are not affected [37].

Several studies have found that acamprosate decreased alcohol drinking or cued responding after a period of chronic ethanolization, when alcohol was presented immediately or after up to a few days of deprivation [36,4244]. Interestingly, this effect was not seen when deprivation lasted 3 weeks [45]. Acamprosate also does not have as pronounced an effect in alcohol-naïve animals [42,44,45[. These findings support Littleton's hypothesis, in that they suggest that the effects of acamprosate may be most prominent close to withdrawal.

Acamprosate does not have abuse potential. Animal experiments suggest that there is little liability for causing self-administration in birds, rodents or primates and there is no generalizability from alcohol [46,47]. In rodents, the direct NMDA antagonist dizopilcine completely substituted for alcohol in a cue-related protocol, but acamprosate did not [47].

Clinical effectiveness of acamprosate

Acamprosate has been demonstrated to be effective in preventing relapse in studies conducted worldwide. Several good-quality systematic reviews and meta-analyses have summarized existing clinical trials, including those by Rosner et al. (Cochrane) [48], National Institute for Clinical Excellence (NICE) CG115 [49], Slattery et al. (Health Technology Board of Scotland) [50], Berglund et al. (Swedish Board) [51] and Bouza et al. (Spanish Agency for Health Technology Assessment) [52], in addition to those by Mann et al. [53]], Kranzler and Van Kirk [54], Mason and Ownby [55], Rosner et al. [56] and Mason and Heyser [57]. Acamprosate increased the chance of abstinence after detoxification by about 15%; Rosner et al. [48] report a relative risk (RR) of 0.86 [95% confidence interval (CI) 0.81–0.91] and NICE CG115 [49] reports RR = 0.83 (95% CI 0.77–0.88), a moderate effect. The ‘number needed to treat’ to result in one more patient becoming abstinent as a result was calculated as 9–11 (e.g. [48,50]). The smaller effect size in later reviews is a consequence of recent studies done in America and Australia where acamprosate was not superior to placebo [5860]. This may reflect the severity of dependence seen in clinical populations in Europe and elsewhere relative to American samples, and some authors have suggested that patients with more severe dependence may benefit more from acamprosate [61]. However, recent meta-analyses [49,62] found no relationship between severity of dependence and response to acamprosate.

In addition to its effect on abstinence, some studies and reviews have found that acamprosate reduces heavy drinking in those patients who do relapse [49,63]. Effects on symptoms associated with alcohol withdrawal have also been reported. Acamprosate has been shown to be effective in decreasing sleep disturbance in withdrawal [64] and during abstinence up to 6 months [65]. It reduced arousal during withdrawal as measured by magnetoencephalography [66]. A slight anxiolytic effect in abstinent alcohol-dependent patients has been reported [67], and it has been found that acamprosate treatment may offset the poorer prognosis conferred by higher anxiety at baseline [68]. Preliminary results of acamprosate adjuvant therapy for treatment of anxiety disorders are promising [69]. Although acamprosate treatment is associated with an increased recovery from depression, this is mediated via the abstinence effect [70].

In observations that replicated preclinical data, acamprosate was less effective when the start of treatment was delayed long after detoxification. A secondary analysis of the COMBINE trial showed that a longer period of pretreatment abstinence predicted a poorer response [71]. The failure of another study to demonstrate efficacy was also attributed to the greater mean duration of abstinence [67]. Although preliminary data have suggested that some drinking outcomes may worsen if acamprosate is started during detoxification, this is still recommended based on the other trial evidence, as well as preclinical evidence of its neuroprotective effect [72]. Certainly, less severe withdrawal symptoms have been reported with acamprosate therapy [60].

The benefits of acamprosate in maintaining abstinence have been shown to persist for 3–12 months after stopping treatment, with a 9% lower risk to return to any drinking in patients who received acamprosate (RR 0.91; 95% CI 0.87–0.96) and a 9% higher continuous abstinence duration (mean difference 8.92; 95% CI 5.08–12.77) [48], compared with those who received placebo. The number needed to treat for prevention of drinking beyond cessation of therapy to post-treatment evaluation was 12.5 (95% CI 9.09–25.00).

There are few predictors of treatment efficacy that might enable us to target acamprosate treatment to patients who might benefit most from it, apart from a goal of abstinence [59] and recent detoxification. A secondary analysis of seven European trials reported that negative family history, late age of onset, anxiety symptoms, severe craving and female gender did not predict response to acamprosate [62]. Psychiatric comorbidity should not limit its use, because it does not worsen symptoms in schizophrenia [73], bipolar affective disorder [74] and depression [70]. However, effectiveness has not been demonstrated in schizophrenia and bipolar affective disorder [70,74].

The usual dose of acamprosate prescribed is 1998 g daily for those over 60 kg and 1332 g for those under 60 kg. A higher dose (3 g day−1) has been reported to be more efficacious [75]. However, a small study compared the tolerability of a 3 g daily dose with a 2 g dose and found that participants reported more nervousness at the higher dose [6]. In general, acamprosate is a well-tolerated drug, with self-limiting diarrhoea being the main side-effect [7].

Acamprosate is frequently described as an ‘anticraving’ drug. Craving is complex and variously conceptualized [76]. Its use clinically may denote a variety of processes, including withdrawal symptoms themselves [77], but also ‘liking, wanting, urges, desires, need, intention or compulsion’ ([76], p. 35). In fact, so problematic was the definition and operationalization of craving that the World Health Organization suggested it should not be used in clinical research into alcohol dependence [78]. It is therefore perhaps not surprising that studies of the effect of acamprosate on craving have not been consistent.

Three clinical trials of acamprosate found a decrease in craving in the acamprosate group relative to placebo at time points within 1 month of withdrawal [30,67] and at 3 months [79]. However, the majority of trials in which craving was measured have found no difference relative to placebo [5860,8085].

Experimental protocols that induce craving have yielded mixed results. treatment with acamprosate for 3 weeks decreased craving after a priming dose of alcohol but not at baseline [4]. Pharmacologically induced craving with yohimbine and meta-chlorophenylpiperazine challenge robustly induced craving, but craving did not diminish in response to acamprosate [86]. Interestingly, head-to-head comparison of naltrexone and acamprosate in reducing cue-induced craving found that naltrexone reduced subjective craving more than acamprosate [87]. In clinical trials where acamprosate and naltrexone were compared with each other and with placebo, naltrexone was reported to reduce craving [58,82] and to reduce drinking when craving was self-rated as high [85], but acamprosate was not. It may be that the questionnaires or visual analog scales do not detect or differentiate that part of craving relating to subclinical withdrawal/negative reinforcement. This is supported by the finding that acamprosate reduced tachycardia associated with craving rather than a subjective measure of craving itself [87].

An alternative conceptualization of the effectiveness of acamprosate in alcohol dependence is that it is substitution therapy, defined as ‘a medication which has one or more of the pharmacological effects of the drug of interest’ [88]. The receptor pharmacology of acamprosate raises this possibility, as does the evidence described above that it ameliorates some symptoms of withdrawal. However, implicit in the concept of substitution is that the subjective effects of the medication are similar to the drug replaced and that the patient experiences dependence on the medication. Acamprosate does not fit this profile; it does not produce alcohol-like intoxication [89], and there are no cases reported of dose escalation and dependence [88].

Acamprosate in other addictions

Following the success of acamprosate in preventing relapse in alcohol dependence, there have been investigations into its utility in treatment of other drug dependencies. Although it reduced the aversive effects of opiate withdrawal in rodents [90], it has not been shown to block opiate reinstatement [91] or sensitization [92]. A single study has reported that acamprosate blocked cue-induced reinstatement of responding to acquire nicotine in rats [93]. Despite early promise that high (>100 mg kg−1), but not lower (30 mg kg−1) doses, of acamprosate diminished cue-and cocaine-induced reinstatement of conditioned place preference [91,94,95], a single randomized controlled trial in cocaine-dependent patients found no significant effect on cocaine use, craving or withdrawal symptoms [96].

Clinical studies of acamprosate have been carried out in pathological gambling and binge-eating disorder. The rationale for the use of acamprosate in pathological gambling was based on early theories that acamprosate acts via GABAA receptors, inhibiting dopaminergic release in response to gambling and thus the experience of reward [97]. A small open-label trial (n = 26) found that the number of gambling episodes and obsessional thinking about gambling decreased [98]. However, a randomized, single-blind trial of acamprosate vs. baclofen (n = 17) found that all participants relapsed within 6 months, and there were no changes from baseline in any subjective measures [97].

Antagonism of the NMDA receptor by acamprosate was suggested to be of value in binge eating, because NMDA receptors in the hypothalamus regulate appetite [99], and other NMDA antagonists, such as memantine, have proved effective in two open-label trials [100,101]. Furthermore, acamprosate treatment is associated with decreased food craving in alcohol-dependent patients [102]. Treatment of binge-eating disorder with acamprosate has been attempted in one, small (n = 20 per group) placebo-controlled study [103]. Although the primary outcome of reduction in bingeing episodes was not different between groups, there was significant decrease in craving as measured by the Yale Brown Obsessive Compulsive Score – Binge Eating and slight weight loss in the active group vs. slight weight gain in the placebo group.

Conclusions

Although acamprosate was initially thought to act at GABAA receptors, the bulk of the literature has subsequently focused on glutamate and NMDA receptors. More is now known about the mechanism by which acamprosate modulates NMDA transmission. Further investigation of other mechanisms of action, for example via glycine and GABAB receptors, is warranted.

The effect of acamprosate on reinstatement of drinking after a period of abstinence in dependent animals has been demonstrated in clinical trials, which have shown increased abstinence and decreased level of drinking after relapse. In rodent studies, this effect is maximal close to withdrawal, and in human studies initiation close to withdrawal provides greater benefit. Acamprosate may also be of benefit as neuroprotection in withdrawal; preclinical and clinical evidence has demonstrated that it decreased the hyperglutamatergic state associated with withdrawal. It has been shown to decrease the behavioural manifestations of withdrawal in rodents and man. Clinical trials suggest that symptoms that are associated with prolonged or subclinical withdrawal, such as anxiety and insomnia, are improved by acamprosate. Taken together, the evidence is consistent with John Littleton's hypothesis that acamprosate targets negative reinforcement craving produced by withdrawal. Unfortunately, most measures of craving in clinical samples do not assess this directly, which may explain why acamprosate has been associated with decreased craving in only a minority of studies in clinical populations. Use of camprosate in other addictive disorders has shown mixed outcomes. Understanding its mechanism of action more fully will allow a more targeted approach to its extension to other uses.

Competing Interests

The authors have completed the Unified Competing Interest form at http://www.icmje.org/coi_disclosure.pdf (available on request from the corresponding author) and declare: no support from any organization for the submitted work; no financial relationships with any organizations that might have an interest in the submitted work in the previous 3 years; no other relationships or activities that could appear to have influenced the submitted work.

N.J.K. is supported by a Wellcome Trust GSK Translational Medicine Training Fellowship.

References

  • 1.Littleton J. Acamprosate in alcohol dependence: how does it work? Addiction. 1995;90:1179–1188. doi: 10.1046/j.1360-0443.1995.90911793.x. [DOI] [PubMed] [Google Scholar]
  • 2.Saivin S, Hulot T, Chabac S, Potgieter A, Durbin P, Houin G. Clinical pharmacokinetics of acamprosate. Clin Pharmacokinet. 1998;35:331–345. doi: 10.2165/00003088-199835050-00001. [DOI] [PubMed] [Google Scholar]
  • 3.Zornoza T, Cano-Cebrian MJ, Hipolito L, Granero L, Polache A. Evidence of a flip-flop phenomenon in acamprosate pharmacokinetics: an in vivo study in rats. Biopharm Drug Dispos. 2006;27:305–311. doi: 10.1002/bdd.513. [DOI] [PubMed] [Google Scholar]
  • 4.Hammarberg A, Jayaram-Lindstrom N, Beck O, Franck J, Reid MS. The effects of acamprosate on alcohol-cue reactivity and alcohol priming in dependent patients: a randomized controlled trial. Psychopharmacology (Berl) 2009;205:53–62. doi: 10.1007/s00213-009-1515-6. [DOI] [PubMed] [Google Scholar]
  • 5.Mason BJ, Goodman AM, Dixon RM, Hameed MH, Hulot T, Wesnes K, Hunter JA, Boyeson MG. A pharmacokinetic and pharmacodynamic drug interaction study of acamprosate and naltrexone. Neuropsychopharmacology. 2002;27:596–606. doi: 10.1016/S0893-133X(02)00368-8. [DOI] [PubMed] [Google Scholar]
  • 6.Johnson BA, O'Malley SS, Ciraulo DA, Roache JD, Chambers RA, Sarid-Segal O, Couper D. Dose-ranging kinetics and behavioral pharmacology of naltrexone and acamprosate, both alone and combined, in alcohol-dependent subjects. J Clin Psychopharmacol. 2003;23:281–293. doi: 10.1097/01.jcp.0000084029.22282.bb. [DOI] [PubMed] [Google Scholar]
  • 7.Rosenthal RN, Gage A, Perhach JL, Goodman AM. Acamprosate: safety and tolerability in the treatment of alcohol dependence. J Addict Med. 2008;2:40–50. doi: 10.1097/ADM.0b013e31816319fd. [DOI] [PubMed] [Google Scholar]
  • 8.al Qatari M, Bouchenafa O, Littleton J. Mechanism of action of acamprosate. Part II. Ethanol dependence modifies effects of acamprosate on NMDA receptor binding in membranes from rat cerebral cortex. Alcohol Clin Exp Res. 1998;22:810–814. [PubMed] [Google Scholar]
  • 9.Naassila M, Hammoumi S, Legrand E, Durbin P, Daoust M. Mechanism of action of acamprosate. Part I. Characterization of spermidine-sensitive acamprosate binding site in rat brain. Alcohol Clin Exp Res. 1998;22:802–809. [PubMed] [Google Scholar]
  • 10.Harris BR, Prendergast MA, Gibson DA, Rogers DT, Blanchard JA, Holley RC, Hart SR, Scotland RL, Foster TC, Pedigo NW, Littleton JM. Acamprosate inhibits the binding and neurotoxic effects of trans-ACPD, suggesting a novel site of action at metabotropic glutamate receptors. Alcohol Clin Exp Res. 2002;26:1779–1793. doi: 10.1097/01.ALC.0000042011.99580.98. [DOI] [PubMed] [Google Scholar]
  • 11.Boismare F, Daoust M, Moore N, Saligaut C, Lhuintre JP, Chretien P, Durlach J. A homotaurine derivative reduces the voluntary intake of ethanol by rats: are cerebral GABA receptors involved? Pharmacol Biochem Behav. 1984;21:787–789. doi: 10.1016/s0091-3057(84)80020-9. [DOI] [PubMed] [Google Scholar]
  • 12.Berton F, Francesconi WG, Madamba SG, Zieglgansberger W, Siggins GR. Acamprosate enhances N-methyl-D-apartate receptor-mediated neurotransmission but inhibits presynaptic GABA(B) receptors in nucleus accumbens neurons. Alcohol Clin Exp Res. 1998;22:183–191. [PubMed] [Google Scholar]
  • 13.Zeise ML, Kasparov S, Capogna M, Zieglgansberger W. Acamprosate (calciumacetylhomotaurinate) decreases postsynaptic potentials in the rat neocortex: possible involvement of excitatory amino acid receptors. Eur J Pharmacol. 1993;231:47–52. doi: 10.1016/0014-2999(93)90682-8. [DOI] [PubMed] [Google Scholar]
  • 14.Reilly MT, Lobo IA, McCracken LM, Borghese CM, Gong D, Horishita T, Harris RA. Effects of acamprosate on neuronal receptors and ion channels expressed in Xenopus oocytes. Alcohol Clin Exp Res. 2008;32:188–196. doi: 10.1111/j.1530-0277.2007.00569.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Pierrefiche O, Daoust M, Naassila M. Biphasic effect of acamprosate on NMDA but not on GABAA receptors in spontaneous rhythmic activity from the isolated neonatal rat respiratory network. Neuropharmacology. 2004;47:35–45. doi: 10.1016/j.neuropharm.2004.03.004. [DOI] [PubMed] [Google Scholar]
  • 16.Daoust M, Legrand E, Gewiss M, Heidbreder C, DeWitte P, Tran G, Durbin P. Acamprosate modulates synaptosomal GABA transmission in chronically alcoholised rats. Pharmacol Biochem Behav. 1992;41:669–674. doi: 10.1016/0091-3057(92)90210-7. [DOI] [PubMed] [Google Scholar]
  • 17.Madamba SG, Schweitzer P, Zieglgansberger W, Siggins GR. Acamprosate (calcium acetylhomotaurinate) enhances the N-methyl-D-aspartate component of excitatory neurotransmission in rat hippocampal CA1 neurons in vitro. Alcohol Clin Exp Res. 1996;20:651–658. doi: 10.1111/j.1530-0277.1996.tb01667.x. [DOI] [PubMed] [Google Scholar]
  • 18.Koob GF, Mason BJ, De Witte P, Littleton J, Siggins GR. Potential neuroprotective effects of acamprosate. Alcohol Clin Exp Res. 2002;26:586–592. [PubMed] [Google Scholar]
  • 19.Popp RL, Lovinger DM. Interaction of acamprosate with ethanol and spermine on NMDA receptors in primary cultured neurons. Eur J Pharmacol. 2000;394:221–231. doi: 10.1016/s0014-2999(00)00195-3. [DOI] [PubMed] [Google Scholar]
  • 20.Rammes G, Mahal B, Putzke J, Parsons C, Spielmanns P, Pestel E, Spanagel R, Zieglgänsberger W, Schadrack J. The anti-craving compound acamprosate acts as a weak NMDA-receptor antagonist, but modulates NMDA-receptor subunit expression similar to memantine and MK-801. Neuropharmacology. 2001;40:749–760. doi: 10.1016/s0028-3908(01)00008-9. [DOI] [PubMed] [Google Scholar]
  • 21.Allgaier C, Franke H, Sobottka H, Scheibler P. Acamprosate inhibits Ca2+ influx mediated by NMDA receptors and voltage-sensitive Ca2+ channels in cultured rat mesencephalic neurones. Naunyn Schmiedebergs Arch Pharmacol. 2000;362:440–443. doi: 10.1007/s002100000285. [DOI] [PubMed] [Google Scholar]
  • 22.Blednov YA, Harris RA. Metabotropic glutamate receptor 5 (mGluR5) regulation of ethanol sedation, dependence and consumption: relationship to acamprosate actions. Int J Neuropsychopharmacol. 2008;11:775–793. doi: 10.1017/S1461145708008584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Chau P, Hoifodt-Lido H, Lof E, Soderpalm B, Ericson M. Glycine receptors in the nucleus accumbens involved in the ethanol intake-reducing effect of acamprosate. Alcohol Clin Exp Res. 2010;34:39–45. doi: 10.1111/j.1530-0277.2009.01063.x. [DOI] [PubMed] [Google Scholar]
  • 24.Erickson CA, Mullett JE, McDougle CJ. Brief report: acamprosate in fragile X syndrome. J Autism Dev Disord. 2010;40:1412–1416. doi: 10.1007/s10803-010-0988-9. [DOI] [PubMed] [Google Scholar]
  • 25.Dahchour A, De Witte P, Bolo N, Nedelec JF, Muzet M, Durbin P, Macher JP. Central effects of acamprosate: part 1. Acamprosate blocks the glutamate increase in the nucleus accumbens microdialysate in ethanol withdrawn rats. Psychiatry Res. 1998;82:107–114. doi: 10.1016/s0925-4927(98)00016-x. [DOI] [PubMed] [Google Scholar]
  • 26.Dahchour A, De Witte P. Ethanol and amino acids in the central nervous system: assessment of the pharmacological actions of acamprosate. Prog Neurobiol. 2000;60:343–362. doi: 10.1016/s0301-0082(99)00031-3. [DOI] [PubMed] [Google Scholar]
  • 27.Zalewska-Kaszubska J, Cwiek W, Dyr W, Czarnecka E. Changes in the beta-endorphin plasma level after repeated treatment with acamprosate in rats selectively bred for high and low alcohol preference. Neurosci Lett. 2005;388:45–48. doi: 10.1016/j.neulet.2005.06.032. [DOI] [PubMed] [Google Scholar]
  • 28.Umhau JC, Momenan R, Schwandt ML, Singley E, Lifshitz M, Doty L, Adams LJ,, Vengeliene V, Spanagel R, Zhang Y, Shen J, George DT, Hommer D, Heilig M. Effect of acamprosate on magnetic resonance spectroscopy measures of central glutamate in detoxified alcohol-dependent individuals: a randomized controlled experimental medicine study. Arch Gen Psychiatry. 2010;67:1069–1077. doi: 10.1001/archgenpsychiatry.2010.125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Kiefer F, Jahn H, Otte C, Nakovics H, Wiedemann K. Effects of treatment with acamprosate on beta-endorphin plasma concentration in humans with high alcohol preference. Neurosci Lett. 2006;404:103–106. doi: 10.1016/j.neulet.2006.05.015. [DOI] [PubMed] [Google Scholar]
  • 30.Hammarberg A, Nylander I, Zhou Q, Jayaram-Lindstrom N, Reid MS, Franck J. The effect of acamprosate on alcohol craving and correlation with hypothalamic pituitary adrenal (HPA) axis hormones and beta-endorphin. Brain Res. 2009;1305(Suppl):S2–6. doi: 10.1016/j.brainres.2009.09.093. [DOI] [PubMed] [Google Scholar]
  • 31.Schule C, Laakmann G, Baghai T, Neukam C, Soyka M. Lack of effects of acamprosate on anterior pituitary hormone secretion in healthy subjects. J Clin Psychopharmacol. 1999;19:387–389. doi: 10.1097/00004714-199908000-00024. [DOI] [PubMed] [Google Scholar]
  • 32.Zimmerman U, Spring K, Koller G, Holsboer F, Soyka M. Hypothalamic-pituitary-adrenal system regulation in recently detoxified alcoholics is not altered by one week of treatment with acamprosate. Pharmacopsychiatry. 2004;37:98–102. doi: 10.1055/s-2004-818986. [DOI] [PubMed] [Google Scholar]
  • 33.Lingford-Hughes A, Watson B, Kalk N, Reid A. Neuropharmacology of addiction and how it informs treatment. Br Med Bull. 2010;96:93–110. doi: 10.1093/bmb/ldq032. [DOI] [PubMed] [Google Scholar]
  • 34.Krystal JH, Staley J, Mason G, Petrakis IL, Kaufman J, Harris RA, Gelernter J, Lappalainen J. Gamma-aminobutyric acid type A receptors and alcohol dependent: intoxication, dependence, vulnerability, and treatment. Arch Gen Psychiatry. 2006;63:957–968. doi: 10.1001/archpsyc.63.9.957. [DOI] [PubMed] [Google Scholar]
  • 35.Tsai G. Glutamatergic neurotransmission in alcohol dependent. J Biomed Sci. 1998;5:309–320. doi: 10.1007/BF02253441. [DOI] [PubMed] [Google Scholar]
  • 36.Gewiss M, Heidbreder C, Opsomer L, Durbin P, De Witte P. Acamprosate and diazepam differentially modulate alcohol-induced behavioural and cortical alterations in rats following chronic inhalation of ethanol vapour. Alcohol Alcohol. 1991;26:129–137. doi: 10.1093/oxfordjournals.alcalc.a045093. [DOI] [PubMed] [Google Scholar]
  • 37.Spanagel R, Putzke J, Stefferl A, Schobitz B, Zieglgansberger W. Acamprosate and alcohol: II. Effects on alcohol withdrawal in the rat. Eur J Pharmacol. 1996;305:45–50. doi: 10.1016/0014-2999(96)00175-6. [DOI] [PubMed] [Google Scholar]
  • 38.Dahchour A, De Witte P. Acamprosate decreases the hypermotility during repeated ethanol withdrawal. Alcohol. 1999;18:77–81. doi: 10.1016/s0741-8329(98)00071-8. [DOI] [PubMed] [Google Scholar]
  • 39.Cole JC, Littleton JM, Little HJ. Acamprosate, but not naltrexone, inhibits conditioned abstinence behaviour associated with repeated ethanol administration and exposure to a plus-maze. Psychopharmacology (Berl) 2000;147:403–411. doi: 10.1007/s002130050009. [DOI] [PubMed] [Google Scholar]
  • 40.Kotlinska J, Bochenski M. The influence of various glutamate receptors antagonists on anxiety-like effect of ethanol withdrawal in a plus-maze test in rats. Eur J Pharmacol. 2008;598:57–63. doi: 10.1016/j.ejphar.2008.09.026. [DOI] [PubMed] [Google Scholar]
  • 41.Farook JM, Krazem A, Lewis B, Morrell DJ, Littleton JM, Barron S. Acamprosate attenuates the handling induced convulsions during alcohol withdrawal in Swiss Webster mice. Physiol Behav. 2008;95:267–270. doi: 10.1016/j.physbeh.2008.05.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Le Magnen J, Tran G, Durlach J, Martin C. Dose-dependent suppression of the high alcohol intake of chronically intoxicated rats by Ca-acetyl homotaurinate. Alcohol. 1987;4:97–102. doi: 10.1016/0741-8329(87)90005-x. [DOI] [PubMed] [Google Scholar]
  • 43.Spanagel R, Holter SM, Allingham K, Landgraf R, Zieglgansberger W. Acamprosate and alcohol: I. Effects on alcohol intake following alcohol deprivation in the rat. Eur J Pharmacol. 1996;305:39–44. doi: 10.1016/0014-2999(96)00174-4. [DOI] [PubMed] [Google Scholar]
  • 44.Heyser CJ, Schulteis G, Durbin P, Koob GF. Chronic acamprosate eliminates the alcohol deprivation effect while having limited effects on baseline responding for ethanol in rats. Neuropsychopharmacology. 1998;18:125–133. doi: 10.1016/S0893-133X(97)00130-9. [DOI] [PubMed] [Google Scholar]
  • 45.Czachowski CL, Delory MJ. Acamprosate and naltrexone treatment effects on ethanol and sucrose seeking and intake in ethanol-dependent and nondependent rats. Psychopharmacology (Berl) 2009;204:335–348. doi: 10.1007/s00213-009-1465-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Grant KA, Woolverton WL. Reinforcing and discriminative stimulus effects of Ca-acetyl homotaurine in animals. Pharmacol Biochem Behav. 1989;32:607–611. doi: 10.1016/0091-3057(89)90005-1. [DOI] [PubMed] [Google Scholar]
  • 47.Spanagel R, Zieglgansberger W, Hundt W. Acamprosate and alcohol: III. Effects on alcohol discrimination in the rat. Eur J Pharmacol. 1996;305:51–56. doi: 10.1016/0014-2999(96)00176-8. [DOI] [PubMed] [Google Scholar]
  • 48.Rosner S, Hackl-Herrwerth A, Leucht S, Lehert P, Vecchi S, Soyka M. Acamprosate for alcohol dependence. Cochrane Database Syst Rev. 2010;(9) doi: 10.1002/14651858.CD004332.pub2. CD004332. [DOI] [PubMed] [Google Scholar]
  • 49.National Institute for Clinical Excellence. Alcohol Dependence and Harmful Alcohol Use. London: NICE; 2011. [Google Scholar]
  • 50.Slattery J, Chick J, Cochrane M, Godfrey C, Kohli H, Macpherson K. Prevention of relapse in alcohol dependence. 2003. Health Technology Assessment Report 3. Health Technology Board for Scotland, Glasgow. Prevention of relapse in alcohol dependence In: 3 HTAR, editor. Glasgow Health Technology Board for Scotland 2003.
  • 51.Berglund M, Thelander S, Jonsson E. Treating Alcohol and Drug Abuse. An Evidence Based Review. KGaAWeinheim: WILEY-VCH Verlag GmbH and Co; 2003. [Google Scholar]
  • 52.Bouza C, Angeles M, Munoz A, Amate JM. Efficacy and safety of naltrexone and acamprosate in the treatment of alcohol dependence: a systematic review. Addiction. 2004;99:811–828. doi: 10.1111/j.1360-0443.2004.00763.x. [DOI] [PubMed] [Google Scholar]
  • 53.Mann K, Lehert P, Morgan MY. The efficacy of acamprosate in the maintenance of abstinence in alcohol-dependent individuals: results of a meta-analysis. Alcohol Clin Exp Res. 2004;28:51–63. doi: 10.1097/01.ALC.0000108656.81563.05. [DOI] [PubMed] [Google Scholar]
  • 54.Kranzler HR, Van Kirk J. Efficacy of naltrexone and acamprosate for alcoholism treatment: a meta-analysis. Alcohol Clin Exp Res. 2001;25:1335–1341. [PubMed] [Google Scholar]
  • 55.Mason BJ, Ownby RL. Acamprosate for the treatment of alcohol dependence: a review of double-blind, placebo-controlled trials. CNS Spectr. 2000;5:58–69. doi: 10.1017/s1092852900012827. [DOI] [PubMed] [Google Scholar]
  • 56.Rosner S, Leucht S, Lehert P, Soyka M. Acamprosate supports abstinence, naltrexone prevents excessive drinking: evidence from a meta-analysis with unreported outcomes. J Psychopharmacol. 2008;22:11–23. doi: 10.1177/0269881107078308. [DOI] [PubMed] [Google Scholar]
  • 57.Mason BJ, Heyser CJ. The neurobiology, clinical efficacy and safety of acamprosate in the treatment of alcohol dependence. Expert Opin Drug Saf. 2010;9:177–188. doi: 10.1517/14740330903512943. [DOI] [PubMed] [Google Scholar]
  • 58.Anton RF, O'Malley SS, Ciraulo DA, Cisler RA, Couper D, Donovan DM, Gastfriend DR, Hosking JD, Johnson BA, LoCastro JS, Longabaugh R, Mason BJ, Mattson ME, Miller WR, Pettinati HM, Randall CL, Swift R, Weiss RD, Williams LD, Zweben A COMBINE Study Research Group. Combined pharmacotherapies and behavioral interventions for alcohol dependence: the COMBINE study: a randomized controlled trial. JAMA. 2006;295:2003–2017. doi: 10.1001/jama.295.17.2003. [DOI] [PubMed] [Google Scholar]
  • 59.Mason BJ, Goodman AM, Chabac S, Lehert P. Effect of oral acamprosate on abstinence in patients with alcohol dependence in a double-blind, placebo-controlled trial: the role of patient motivation. J Psychiatr Res. 2006;40:383–393. doi: 10.1016/j.jpsychires.2006.02.002. [DOI] [PubMed] [Google Scholar]
  • 60.Morley KC, Teesson M, Reid SC, Sannibale C, Thomson C, Phung N, Weltman M, Bell JR, Richardson K, Haber PS. Naltrexone versus acamprosate in the treatment of alcohol dependence: A multi-centre, randomized, double-blind, placebo-controlled trial. Addiction. 2006;101:1451–1462. doi: 10.1111/j.1360-0443.2006.01555.x. [DOI] [PubMed] [Google Scholar]
  • 61.Garbutt JC, Kampov-Polevoy AB, Gallop R, Kalka-Juhl L, Flannery BA. Efficacy and safety of baclofen for alcohol dependence: a randomized, double-blind, placebo-controlled trial. Alcohol Clin Exp Res. 2010;34:1849–1857. doi: 10.1111/j.1530-0277.2010.01273.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Verheul R, Lehert P, Geerlings PJ, Koeter MW, van den Brink W. Predictors of acamprosate efficacy: results from a pooled analysis of seven European trials including 1485 alcohol-dependent patients. Psychopharmacology (Berl) 2005;178:167–173. doi: 10.1007/s00213-004-1991-7. [DOI] [PubMed] [Google Scholar]
  • 63.Chick J, Lehert P, Landron F. Does acamprosate improve reduction of drinking as well as aiding abstinence? J Psychopharmacol. 2003;17:397–402. doi: 10.1177/0269881103174017. [DOI] [PubMed] [Google Scholar]
  • 64.Staner L, Boeijinga P, Danel T, Gendre I, Muzet M, Landron F, Luthringer R. Effects of acamprosate on sleep during alcohol withdrawal: A double-blind placebo-controlled polysomnographic study in alcohol-dependent subjects. Alcohol Clin Exp Res. 2006;30:1492–1499. doi: 10.1111/j.1530-0277.2006.00180.x. [DOI] [PubMed] [Google Scholar]
  • 65.Perney P, Lehert P, Mason BJ. Sleep disturbance in alcoholism: proposal of a simple measurement, and results from a 24-week randomized controlled study of alcohol-dependent patients assessing acamprosate efficacy. Alcohol Alcohol. 2012;47:133–139. doi: 10.1093/alcalc/agr160. [DOI] [PubMed] [Google Scholar]
  • 66.Boeijinga PH, Parot P, Soufflet L, Landron F, Danel T, Gendre I, Muzet M, Demazières A, Luthringer R. Pharmacodynamic effects of acamprosate on markers of cerebral function in alcohol-dependent subjects administered as pretreatment and during alcohol abstinence. Neuropsychobiology. 2004;50:71–77. doi: 10.1159/000077944. [DOI] [PubMed] [Google Scholar]
  • 67.Chick J, Howlett H, Morgan MY, Ritson B. United Kingdom Multicentre Acamprosate Study (UKMAS): a 6-month prospective study of acamprosate versus placebo in preventing relapse after withdrawal from alcohol. Alcohol Alcohol. 2000;35:176–187. doi: 10.1093/alcalc/35.2.176. [DOI] [PubMed] [Google Scholar]
  • 68.Mason BJ, Lehert P. The effects of current subsyndromal psychiatric symptoms or past psychopathology on alcohol dependence treatment outcomes and acamprosate efficacy. Am J Addict. 2010;19:147–154. doi: 10.1111/j.1521-0391.2009.00013.x. [DOI] [PubMed] [Google Scholar]
  • 69.Schwartz TL, Siddiqui UA, Raza S, Costello A. Acamprosate calcium as augmentation therapy for anxiety disorders. Ann Pharmacother. 2010;44:1930–1932. doi: 10.1345/aph.1P353. [DOI] [PubMed] [Google Scholar]
  • 70.Lejoyeux M, Lehert P. Alcohol-use disorders and depression: results from individual patient data meta-analysis of the acamprosate-controlled studies. Alcohol Alcohol. 2011;46:61–67. doi: 10.1093/alcalc/agq077. [DOI] [PubMed] [Google Scholar]
  • 71.Gueorguieva R, Wu R, Donovan D, Rounsaville BJ, Couper D, Krystal JH, Krystal JH, O'Malley SS. Baseline trajectories of drinking moderate acamprosate and naltrexone effects in the COMBINE study. Alcohol Clin Exp Res. 2011;35:523–531. doi: 10.1111/j.1530-0277.2010.01369.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Lingford-Hughes AR, Welch S, Peters L, Nutt DJ. Evidence-based guidelines for the pharmacological management of substance misuse, addiction and comorbidity: recommendations from the British Association for Psychopharmacology. J Psychopharmacol. 2012;26:1–54. doi: 10.1177/026988110401800321. [DOI] [PubMed] [Google Scholar]
  • 73.Ralevski E, O'Brien E, Jane JS, Dean E, Dwan R, Petrakis I. Effects of acamprosate on cognition in a treatment study of patients with schizophrenia spectrum disorders and comorbid alcohol dependence. J Nerv Ment Dis. 2010;199:499–505. doi: 10.1097/NMD.0b013e3182214297. [DOI] [PubMed] [Google Scholar]
  • 74.Tolliver BK, Desantis SM, Brown DG, Prisciandaro JJ, Brady KT. A randomized, double-blind, placebo-controlled clinical trial of acamprosate in alcohol-dependent individuals with bipolar disorder: a preliminary report. Bipolar Disord. 2012;14:54–63. doi: 10.1111/j.1399-5618.2011.00973.x. [DOI] [PubMed] [Google Scholar]
  • 75.Mason BJ. Acamprosate and naltrexone treatment for alcohol dependence: an evidence-based risk-benefits assessment. Eur Neuropsychopharmacol. 2003;13:469–475. doi: 10.1016/j.euroneuro.2003.08.009. [DOI] [PubMed] [Google Scholar]
  • 76.Drummond DC. Theories of drug craving, ancient and modern. Addiction. 2001;96:33–46. doi: 10.1046/j.1360-0443.2001.961333.x. [DOI] [PubMed] [Google Scholar]
  • 77.Kozlowski LT, Wilkinson DA. Use and misuse of the concept of craving by alcohol, tobacco, and drug researchers. Br J Addict. 1987;82:31–45. doi: 10.1111/j.1360-0443.1987.tb01430.x. [DOI] [PubMed] [Google Scholar]
  • 78.World Health Organisation. Alcohol and Alcoholism: Report of An Expert Committee. Geneva: WHO; 1955. p. 94. [Google Scholar]
  • 79.Paille FM, Guelfi JD, Perkins AC, Royer RJ, Steru L, Parot P. Double-blind randomized multicentre trial of acamprosate in maintaining abstinence from alcohol. Alcohol Alcohol. 1995;30:239–247. [PubMed] [Google Scholar]
  • 80.Sass H, Soyka M, Mann K, Zieglgansberger W. Relapse prevention by acamprosate. Results from a placebo-controlled study on alcohol dependence. Arch Gen Psychiatry. 1996;53:673–680. doi: 10.1001/archpsyc.1996.01830080023006. [DOI] [PubMed] [Google Scholar]
  • 81.Poldrugo F. Acamprosate treatment in a long-term community-based alcohol rehabilitation programme. Addiction. 1997;92:1537–1546. [PubMed] [Google Scholar]
  • 82.Rubio G, Jimenez-Arriero MA, Ponce G, Palomo T. Naltrexone versus acamprosate: one year follow-up of alcohol dependence treatment. Alcohol Alcohol. 2001;36:419–425. doi: 10.1093/alcalc/36.5.419. [DOI] [PubMed] [Google Scholar]
  • 83.Tempesta E, Janiri L, Bignamini A, Chabac S, Potgieter A. Acamprosate and relapse prevention in the treatment of alcohol dependence: a placebo-controlled study. Alcohol Alcohol. 2000;35:202–209. doi: 10.1093/alcalc/35.2.202. [DOI] [PubMed] [Google Scholar]
  • 84.Namkoong K, Lee BO, Lee PG, Choi MJ, Lee E. Acamprosate in Korean alcohol-dependent patients: a multi-centre, randomized, double-blind, placebo-controlled study. Alcohol Alcohol. 2003;38:135–141. doi: 10.1093/alcalc/agg038. [DOI] [PubMed] [Google Scholar]
  • 85.Richardson K, Baillie A, Reid S, Morley K, Teesson M, Sannibale C, Weltman M, Haber P. Do acamprosate or naltrexone have an effect on daily drinking by reducing craving for alcohol? Addiction. 2008;103:953–959. doi: 10.1111/j.1360-0443.2008.02215.x. [DOI] [PubMed] [Google Scholar]
  • 86.Umhau JC, Schwandt ML, Usala J, Geyer C, Singley E, George DT, Heilig M. Pharmacologically induced alcohol craving in treatment seeking alcoholics correlates with alcoholism severity, but is insensitive to acamprosate. Neuropsychopharmacology. 2011;36:1178–1186. doi: 10.1038/npp.2010.253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Ooteman W, Koeter MW, Verheul R, Schippers GM, van denBrink W. The effect of naltrexone and acamprosate on cue-induced craving, autonomic nervous system and neuroendocrine reactions to alcohol-related cues in alcoholics. Eur Neuropsychopharmacol. 2007;17:558–566. doi: 10.1016/j.euroneuro.2007.02.012. [DOI] [PubMed] [Google Scholar]
  • 88.Nutt DJ, Lingford-Hughes A, Chick J. Through a glass darkly: can we improve clarity about mechanism and aims of medications in drug and alcohol treatments? J Psychopharmacol. 2012;26:199–204. doi: 10.1177/0269881111410899. [DOI] [PubMed] [Google Scholar]
  • 89.Brasser SM, McCaul ME, Houtsmuller EJ. Alcohol effects during acamprosate treatment: a dose-response study in humans. Alcohol Clin Exp Res. 2004;28:1074–1083. doi: 10.1097/01.alc.0000130802.07692.29. [DOI] [PubMed] [Google Scholar]
  • 90.Kratzer U, Schmidt WJ. The anti-craving drug acamprosate inhibits the conditioned place aversion induced by naloxone-precipitated morphine withdrawal in rats. Neurosci Lett. 1998;252:53–56. doi: 10.1016/s0304-3940(98)00550-3. [DOI] [PubMed] [Google Scholar]
  • 91.McGeehan AJ, Olive MF. The anti-relapse compound acamprosate inhibits the development of a conditioned place preference to ethanol and cocaine but not morphine. Br J Pharmacol. 2003;138:9–12. doi: 10.1038/sj.bjp.0705059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kratzer U, Spanagel R, Schmidt WJ. The effect of acamprosate on the development of morphine-induced behavioral sensitization in rats. Behav Pharmacol. 2003;14:351–356. doi: 10.1097/01.fbp.0000080417.18561.77. [DOI] [PubMed] [Google Scholar]
  • 93.Pechnick RN, Manalo CM, Lacayo LM, Vit JP, Bholat Y, Spivak I, Reyes KC, Farrokhi C. Acamprosate attenuates cue-induced reinstatement of nicotine-seeking behavior in rats. Behav Pharmacol. 2011;22:222–227. doi: 10.1097/FBP.0b013e328345f72c. [DOI] [PubMed] [Google Scholar]
  • 94.McGeehan AJ, Olive MF. Attenuation of cocaine-induced reinstatement of cocaine conditioned place preference by acamprosate. Behav Pharmacol. 2006;17:363–367. doi: 10.1097/01.fbp.0000224384.01863.5f. [DOI] [PubMed] [Google Scholar]
  • 95.Bowers MS, Chen BT, Chou JK, Osborne MP, Gass JT, See RE, Bonci A, Janak PH, Olive MF. Acamprosate attenuates cocaine-and cue-induced reinstatement of cocaine-seeking behavior in rats. Psychopharmacology (Berl) 2007;195:397–406. doi: 10.1007/s00213-007-0904-y. [DOI] [PubMed] [Google Scholar]
  • 96.Kampman KM, Dackis C, Pettinati HM, Lynch KG, Sparkman T, O'Brien CP. A double-blind, placebo-controlled pilot trial of acamprosate for the treatment of cocaine dependence. Addict Behav. 2011;36:217–221. doi: 10.1016/j.addbeh.2010.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Dannon PN, Rosenberg O, Schoenfeld N, Kotler M. Acamprosate and Baclofen were Not Effective in the Treatment of Pathological Gambling: Preliminary Blind Rater Comparison Study. Front Psychiatry. 2011;2:1–4. doi: 10.3389/fpsyt.2011.00033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Black DW, McNeilly DP, Burke WJ, Shaw MC, Allen J. An open-label trial of acamprosate in the treatment of pathological gambling. Ann Clin Psychiatry. 2011;23:250–256. [PubMed] [Google Scholar]
  • 99.Stanley BG, Willett VL, 3rd, Donias HW, Dee MG, 2nd, Duva MA. Lateral hypothalamic NMDA receptors and glutamate as physiological mediators of eating and weight control. Am J Physiol. 1996;270(2 Pt 2):R443–449. doi: 10.1152/ajpregu.1996.270.2.R443. [DOI] [PubMed] [Google Scholar]
  • 100.Hermanussen M, Tresguerres JA. A new anti-obesity drug treatment: first clinical evidence that, antagonising glutamate-gated Ca2+ ion channels with memantine normalises binge-eating disorders. Econ Hum Biol. 2005;3:329–337. doi: 10.1016/j.ehb.2005.04.001. [DOI] [PubMed] [Google Scholar]
  • 101.Brennan BP, Roberts JL, Fogarty KV, Reynolds KA, Jonas JM, Hudson JI. Memantine in the treatment of binge eating disorder: an open-label, prospective trial. Int J Eat Disord. 2008;41:520–526. doi: 10.1002/eat.20541. [DOI] [PubMed] [Google Scholar]
  • 102.Han DH, Lyool IK, Sung YH, Lee SH, Renshaw PF. The effect of acamprosate on alcohol and food craving in patients with alcohol dependence. Drug Alcohol Depend. 2008;93:279–283. doi: 10.1016/j.drugalcdep.2007.09.014. [DOI] [PubMed] [Google Scholar]
  • 103.McElroy SL, Guerdjikova AI, Winstanley EL, O'Melia AM, Mori N, McCoy J, Keck PE, Jr, Hudson JI. Acamprosate in the treatment of binge eating disorder: a placebo-controlled trial. Int J Eat Disord. 2011;44:81–90. doi: 10.1002/eat.20876. [DOI] [PubMed] [Google Scholar]

Articles from British Journal of Clinical Pharmacology are provided here courtesy of British Pharmacological Society

RESOURCES