Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Jan 1.
Published in final edited form as: Curr Pharm Des. 2008;14(15):1512–1524. doi: 10.2174/138161208784480090

Potential therapeutic interest of adenosine A2A receptors in psychiatric disorders

Rodrigo A Cunha 1, Sergi Ferré 2, Jean-Marie Vaugeois 3, Jiang-Fan Chen 4
PMCID: PMC2423946  NIHMSID: NIHMS51865  PMID: 18537674

Abstract

The interest on targeting adenosine A2A receptors in the realm of psychiatric diseases first arose based on its tight physical and functional interaction with dopamine D2 receptors. However, the role of central A2A receptors is now viewed as much broader than just controlling D2 receptor function. Thus, there is currently a major interest in the ability of A2A receptors to control synaptic plasticity at glutamatergic synapses. This is due to a combined ability of A2A receptors to facilitate the release of glutamate and the activation of NMDA. Therefore, A2A receptors are now conceived as a normalizing device promoting adequate adaptive responses in neuronal circuits, a role similar to that fulfilled, in essence, by dopamine. This makes A2A receptors a particularly attractive target to manage psychiatric disorders since adenosine may act as go-between glutamate and dopamine, two of the key players in mood processing. Furthermore, A2A receptors also control glia function and brain metabolic adaptation, two other emerging mechanisms to understand abnormal processing of mood, and A2A receptors are an important player in controlling the demise of neurodegeneration, considered an amplificatory loop in psychiatric disorders. Current data only provide an indirect confirmation of this putative role of A2A receptors, based on the effects of caffeine (an antagonist of both A1 and A2A receptors) in psychiatric disorders. However, the introduction of A2A receptors in clinics as anti-parkinsonian agents is hoped to bolster our knowledge on the role of A2A receptors in mood disorders in the near future.

Keywords: adenosine, A2A receptor, caffeine, mood disorders, psychiatric diseases, anxiety, depression, schizophrenia, attention deficit hyperactivity disorder, ADHD

INTRODUCTION

Psychiatric disorders are currently defined on the basis of behavioural modifications found in patients. Behavioural analysis essentially provides trends suggesting modified behavioural patterns in comparison with a standardised population, which in itself display intra- and inter-subject heterogeneity. There is currently no clear bio-marker to support the modified behavioural patterns. This might be one of the reasons justifying the difficulty in categorising psychiatric disorders, in spite of the tremendous effort in the refinement of neuropsychological tests.

This reality also makes it difficult to appreciate the relevance of novel molecular targets to develop drugs aimed at managing psychiatric conditions. In fact, the decision on pursuing a given molecular target to develop novel therapeutics is expected to be based on a strong scientific rational. This normally derives from the pathological changes that are characteristic of the disease conditions being targeted. In the case of brain disorders, one should ideally identify what brain areas are primarily affected and what are the main biochemical and/or neurochemical traits pathognomonic of the disease. For instance, it would be of great help deciding if the disease is mostly associated with neuronal or glial deficit. In case it would be mostly neuronal, one could seek for the brain circuits primarily affected, and the neurochemical systems suffering the most significant imbalance; if a glial deficit would be evident, then one could attempt defining if the disease results from a metabolic drift or if neuroinflammation deregulation plays a role. Finally, the issue of neurogenesis defaults as a possible cause of disease should also be considered. It is this general information that ultimately provides the scientific rationale to select any particular molecular target to develop novel therapeutic strategies.

In the case of psychiatric disorders, it is currently not possible to apply any solid anatomical or neurochemical rationale to sustain pursuing any particular molecular target for the development of novel drug-based therapeutic strategy. In fact, none of the questions listed above have received a clear answer in the case of the most common psychiatric disorders. Taking as examples the case of depressive disorders (the plural reflecting the idea that they are multiple defined clinical entities), the brain areas affected are rather broad and too many biochemical and/or neurochemical (or morphological) traits have been reported to allow any of them to be considered pathognomonic of these ‘diseases’ [15]. Different groups place different emphasis on whether ‘depression’ is primarily due to neuronal or glial modifications [6,7]. Accordingly, there is no clear definition of particular brain circuits affected in these conditions, nor there is any agreement on whether these conditions are due to metabolic [810] or neuroinflammatory deregulations [11,12]. Finally the currently holly grail of therapeutics (neurogenesis) actually seems to be a part in all physiological and pathological processes in the brain [13,14], making it difficult to anticipate how this can be manipulated as a therapeutic strategy.

Without a clear rationale to discuss the validity of considering any particular molecular target as a promising candidate to develop novel drugs to manage psychiatric conditions, one is left with the evaluation of the efficacy of novel drugs in alleviating the behavioural symptoms that are characteristic of these diseases. The development of drugs is normally carried out in a safer and faster manner using animal models of disease. And this constitutes the second major hurl to test the interest of potentially novel drugs to manage psychiatric disorders. In fact, there is currently no single animal model that satisfactory mimics the most common behavioural changes found in psychiatric disorders [1517]. There are obviously animal models that replicate particular behavioural changes (but only a limited set) and some animal behavioural tests providing a reasonable predictability of the efficiency of some (but not all) of the drugs currently used to alleviate the symptoms of psychiatric disorders [18,19].

The recognition of our current limitations in exploring novel targets to develop new drug-based therapeutic strategies to manage psychiatric disorders should be kept in mind when evaluating the subsequently presented evidence suggesting the possible interest of adenosine A2A receptors.

PHYSIOLOGICAL ROLE(S) OF ADENOSINE A2A RECEPTORS IN THE BRAIN

There are several reviews dealing with the localization and role of A2A receptors in the brain [2023]. This short overview is just supposed to recapitulate some features of central A2A receptors that might be relevant to the putative interest of this receptor in the realm of psychiatric disorders.

Until the beginning of this century, there was a general consensus that central A2A receptors were confined to the basal ganglia, where they played a role in the control of signal processing in medium spiny neurons [2427]. In fact, this particular pool of A2A receptors is by far the most abundant in the mammalian brain, but this should not underscore the fact that A2A receptors have a much broader distribution in different brain areas and in different cell types, albeit with a considerably lower density. These A2A receptors in medium spiny neurons have been established to be determinant for the control of motor function, since their selective genetic elimination abrogates the ability of A2A receptor to control motor function [28], probably the most evident behavioural effect caused by A2A receptor ligands [23,2931].

How these A2A receptors located in medium spiny neurons act to control motor function is still an open issue (reviewed in [31]). There is a predominant trend arguing that the main action of these striatal A2A receptors is the control of dopaminergic signalling, that plays a key role in striatal signal processing and thus in motor control [32]. In particular the pioneering work at the Karolinska Institute (reviewed in Ferré et al., present issue) clearly substantiated a tight interaction between A2A and dopamine D2 receptor signalling. However, it is also clear that A2A receptors can control motor function in the absence of dopaminergic signalling [33,34]. This indicates that even striatal A2A receptors work through dopamine-independent mechanisms to impact on brain function. In fact, A2A receptors have been localized presynaptically in a majority of glutamatergic nerve terminals, where they form heteromers with A1 receptors and where they play an important facilitatory role of cortico-striatal glutamatergic neurotransmission [35].

The concept of dopamine-independent effects of A2A receptor function is particularly relevant in the case of extra-striatal A2A receptors, where dopaminergic signalling is far less intense. The most compelling evidence come from the recent study using the brain-region specific A2A receptor knockout models in which A2A receptor was selectively deleted either in striatal neurons (striatum A2A KO) or entire forebrain neurons (including striatum, cerebral cortex and hippocampus, forebrain A2A KO) [3638]. Using these novel knockout models, we recently showed that cocaine-induced psychomotor activity is enhanced in striatum A2A KO mice, but attenuated in forebrain A2A KO mice; urthermore, selective inactivation of A2A receptor in extra-striatal cells by administering the A2A receptor antagonist KW6002 to striatum A2A KO mice attenuated cocaine effects, rather than enhanced cocaine effects by administering KW6002 into wild-type mice [39]. These results identify a critical role of A2A receptors in extra-striatal neurons in providing a prominent excitatory effect on psychomotor activity [39]. Theprecise localization of these extra-striatal A2A receptors involved in psychomotor is not clear yet, but several studies have found that these extra-striatal A2A receptors are mostly synaptically-located in contrast to the most abundant striatal A2A receptors [40]. In particular, extra-striatal A2A receptors are located in glutamatergic synapses [41]. It is important to point out that those extra-striatal A2A receptors also include the A2A receptors localized in striatal glutamatergic terminals [35] (thus, the term extra-striatal can be a bit misleading). These A2A receptors control both the release of glutamate [35,42,43] as well as NMDA receptors [44]. Interestingly, these receptors do not seem to be activated by ambient levels of adenosine [4446]. Instead, they are selectively recruited upon high frequency trains of afferent stimulation that are normally used to trigger synaptic plasticity phenomena [44]. This is due to the fact that A2A receptors seem to be selectively activated by a pool of adenosine formed upon the extracellular catabolism of ATP [44,47], which is mainly released upon higher frequencies of nerve stimulation [48]. This engagement of A2A receptors selectively upon high frequency trains of stimulation designed to trigger plastic changes in excitatory synapses has lead to the proposal that the adenosine system would help defining salience of information in excitatory circuits through a combined action of A2A receptors, as an ancillary system of NMDA receptors, in synapses engaged in plastic changes, together with the action of inhibitory A1 receptors (activated through astrocytic-mediated heterosynaptic depression) in non-stimulated synapses [21]. Hence A2A receptors would play a selective role in controlling plastic changes in brain circuits, defining the threshold for induction of plastic changes in excitatory synapses.

Other possible physiological functions potentially controlled by A2A receptors are also worth considering, although the weight of evidence in their support is currently weaker. One aspect that merits further investigation is the possible ability of A2A receptors to control inhibitory transmission in brain circuits. Neurochemical findings showed that A2A receptors controlled the evoked release of GABA from different preparations [49,50], but this has only received a direct electrophysiological support in the adult brain in the case of collateral projection between medium spiny neurons [51] and in their projection to the pallidus [52]. This topic is of particular relevance given the importance of long-distance interneurons and local interneurons in the definition of cortical excitability [53]. The interest on this subject is strengthened by the recent observation that adenosine receptor blockade following caffeine administration seems to mainly affect inhibitory rather than excitatory transmission in the Human cortex [54]. Another potential role of A2A receptors in physiological conditions is as coordinator of metabolic activity in brain tissue. Thus, adenosine has long been recognised as a key paracrine modulator in different mammalian tissue, being responsible for function such as cardiac dromotropism, tuberulo-glomerular filtration control, post-prandial vasodilatation and control of excessive immune/inflammatory reactivity [55]. In fact, ATP (one of the most abundant intracellular molecules) and adenosine are released from stressed cells (either suffering insults or upon work overload) and this extracellular adenosine acts on both A1 and A2A receptors to prompt adaptation and/or restore homeostasis [56,57]. In brain tissues, A2A receptors control capillary vasodilatation [58], the uptake of excitatory amino acids by astrocytes and the pattern of metabolism in astrocytes [59]. This is expected to have a dramatic impact both on the availability and use of metabolic resources that are fundamental to the optimal performance of brain circuits, but the true contribution of A2A receptors for brain metabolism still needs to be thoughtfully tested.

ROLE OF ADENOSINE A2A RECEPTORS IN THE CONTROL OF NEURODEGENERATION

The impact of A2A receptors in the control of neuronal damage was first proposed by John Phillis in a model of cerebral ischemic injury [60]. It was later confirmed that either the pharmacological blockade or the genetic elimination of A2A receptors conferred a robust neuroprotection in animal models of brain ischemia [61,62]. This was later extended to a variety of situations that had in common the deleterious impact of chronic noxious insults to adult brain tissue (reviewed in [20,57]), such as glutamate excitotoxicity [6365], free radical toxicity [66], epilepsy [6769], MPTP toxicity [7072], 6-hydroxydopamine toxicity [70,71], 3-nitropropionic acid toxicity [37,73,74] or β-amyloid toxicity [75,76]. Interestingly, the neuroprotection afforded by A2A receptor blockade is most evident in cortical areas (reviewed in [57]), where the density of A2A receptors is nearly 20 times lower than in the striatum [77]. It is important to note that the neuroprotective effect of A2A receptor antagonists in general correlates with their ability to improve cognitive behaviour in animal models of neurological disorders [20,57,78]. Consequently, A2A receptor activity in brain may achieve the modulation of cognitive function, particularly those associated with degenerative disorders (such as Parkinson’s disease, Huntington’s disease and Alzheimer’s disease), through its control of neuronal cell death.

The mechanism underlying this ability of A2A receptors to impact on brain tissue damage is still a matter of hot debate. The use of tissue-specific transgenic mice fostered by our group in Boston University School of Medicine, indicated that non-neuronal A2A receptors were responsible for the control of brain tissue damage; in ischemic models or models of 3-nitropropionic acid-induced toxicity, it was concluded that the key role was played by A2A receptors from bone marrow-derived cells [37,79], whereas in MPTP-induced neurotoxicity A2A receptors in glial cells were the ones that played the key role in controlling brain tissue damage [28]. This is in agreement with the localization of A2A receptors in microglia cells and their ability to control microglia activation and burst of neuroinflammation [80]. However, there is also robust evidence showing that neuronal A2A receptors can also control the demise of neuronal damage. This was shown in the case of cultured neurons (virtually devoid of microglia or inflammatory cells), where A2A receptor blockade abrogated either β-amyloid- [81] or staurosporine-induced neurotoxicity [82] through a control of mitochondria membrane potential and release of pro-apoptotic factors. These stimuli caused an initial synaptic damage that later evolved into overt loss of neuronal viability, in light of the particular synaptic localization of cortical A2A receptors and with the wide spreading idea that chronic neurodegenerative diseases begin with synaptic dysfunctions that later evolve into different demises of neurodegeneration [8385]. In agreement with this role of synaptic A2A receptors in the control of brain tissue damage is the observation that A2A receptor antagonists prevented restraint stress-induced synaptic damage in the CA3 area of the rat hippocampus without any apparent involvement of changes in inflammatory-related cells [86]. Clearly, this existence of multiple and apparently conflicting hypothesis illustrate how little we actually know about the different possible demises of brain tissue damage as well as of how little we know on the biology of A2A receptors.

A consensual idea would be to propose that there might be a successive participation of A2A receptors located in different cells types according to the duration and/or intensity of noxious brain insult: with mild noxious insults, there might be a main role of synaptic A2A receptors; with more prolonged noxious stimuli, microglia A2A receptors would play a predominant role, in view of the importance of microglia in the amplification of early brain damage [8789]; finally, with more severe damage, causing loss of preservation of the blood-brain barrier, it might be that A2A receptors in inflammatory cells invading the brain parenchyma play the more pronounced role. Clearly, this is a hypothetic scenario that still needs experimental confirmation.

A final topic that deserves consideration is the transducing mechanisms operated by A2A receptors to fulfil their physiological role(s) and to impact on brain tissue damage. There is general agreement in the field that the transducing system operated by adenosine A2A receptors is through the adenylate cyclase/cAMP/protein kinase A pathway [90]. This has received direct experimental confirmation in heterologous expression system (where this was the only pathway that was investigated) and in striatal medium spiny neurons [9192]. However, it is now clear that A2A receptors can couple to different transducing pathways (reviewed in [23,57]), being a prototypical example of a pleiotropic receptor. At least for its impact on neuroprotection, it is clear and evident that A2A receptors do not act through the cAMP pathway: in fact, it is well known that bursting the cAMP pathway affords neuroprotection [93,94]; in contrast, it is the blockade of A2A receptors (which would trigger but rather prevent accumulations of cAMP) that actually confers neuroprotection. The clarification of the transducing pathways operated by A2A receptors is an issue of particular relevance since “normalisation of signaling” through manipulating A2A receptors is a potential important issue in the realm of psychiatric disorders.

ADENOSINE AND MOOD DISORDERS

Mood disorders are one of the greatest burdens of disease in Europe and the development of effective strategies to manage these conditions should represent a major socio-economic priority [9597]. The interest in the role of adenosine in mood disorders stems from three concurrent lines of research: first, there is evidence that the consumption of coffee, and particularly of caffeine (an adenosine receptor antagonists, as discussed below) might modify the mood profile both of volunteers as well as of psychiatric patients; secondly, there is evidence that different therapeutic strategies used to control mood disorders cause effects related to the adenosine modulation system; thirdly, there is evidence from animal models that the manipulation of adenosine receptors modifies behavioural responses considered relevant for mood function in Humans. These first two lines of evidence will be discussed in parallel, whereas the last one will be separately discussed since it is the only one that allows directly relating A2A receptors with mood disorders (until data from the use of A2A receptor ligands in Humans becomes publicly available).

Several studies in Humans have explored the relation between coffee intake and the mood changes. These studies are likely to be relevant to the understanding of the putative role of the adenosine modulation system in the control of mood for two reasons: first because it is becoming evident that most of the effects of caffeine on brain related functions are mostly due to the effects of caffeine, since they are not mimicked by decaffeinated coffee or other drinks such as fruit juice (reviewed in [98]); secondly, the only known molecular target of caffeine at physiological (i.e. nontoxic) doses are the A1 and A2A adenosine receptors, where caffeine acts as a competitive antagonist [99,100]. The consumption of coffee is well documented to increase alertness (reviewed in [98,101]) and there is a trend to consider that caffeine improves performance and cognition, especially in situations decreasing performance of cognition (reviewed in [20,57,78]). There is also a general perception that caffeine consumption may render individuals more anxious. Actually, large consumption of coffee (or caffeine) has been argued to trigger a constellation of behavioural modifications that has led to coining the term ‘caffeinism’ [102104]. In this situation, there are both anxiety disorders as well as greater incidence of depressive-like conditions [103,105]. Another situation where there is a strong link between caffeine intake and modifications of mood is upon withdrawal of caffeine [106,107]. Apart from headache, fatigue and decreased alertness [108109], withdrawal from regular consumption of caffeine triggers a variety of anxiety-like symptoms, such as irritability, sleepiness, dysphoria, nervousness or restlessness [106,107,110112]. It is interesting to note that some of these same withdrawal symptoms are similar to those described to occur upon ‘caffeinism’. This leads to two inter-twinned ideas that should be kept in mind when evaluating the putative role(s) of adenosine and its receptors in the control of mood. The first idea is that adenosine (and in an inverse manner caffeine) act on two receptors with globally opposite function, namely inhibitory A1 and facilitatory A2A receptors. Hence, it is possible that different concentrations (or doses) of caffeine and adenosine may cause opposite effects operated by different receptors. The second idea is a re-phrasal of the previous idea, i.e. that the adenosine neuromodulation system should be viewed as a paracrine system designed to maintain homeostasis or promote adaptation of neuronal systems. This means that the fundamental role of this adenosine modulation system is to narrow the window of functioning of biological systems, curtailing its edges of extremes of functioning. Adhering to these ideas will make it obvious that two much or too little adenosine in a system will cause its failure to properly adapt to its environment. This might be a possible underlying cause to explain the similarity between withdrawn of caffeine and ‘caffeinism’

A second line of evidence that is suggestive of a role of adenosine receptors in the control of mood is the observations that different therapeutic strategies used to control mood disorders have effects related to the adenosine modulation system [113]. In fact, both electroconvulsive therapy and sleep deprivation are two types of treatments of mood disorders, both of which causing short term and long term adaptations of the adenosine neuromodulation system. Thus, there are short term adaptive neuronal responses that are operated through inhibitory A1 receptors, namely in terms of the slow wave sleep [114] and cerebral metabolic activity [115,116]. There are also more long term adaptive changes, such as up-regulation of A1 receptors [117,119] and possibly of A2A receptors (reviewed in [57]) the former being a strong candidate to mediate the reduction of cerebral blood flow [116,120122], which is observed after these treatments. It should be made clear that at this stage there is a tentative parallel between the effects operated by these mood disorder treatments and the adenosine modulation system in the brain, but it still remain to be directly shown that the mood beneficial effects of these treatments is hampered by manipulation of adenosine receptors.

ADENOSINE A2A RECEPTORS AND ANXIETY

The role of adenosine A2A receptors in anxiety is still to be defined. In fact, whereas higher doses of caffeine tend to increase [103,123127] and lower doses of caffeine tend to reduce anxiety levels in Humans [128,129], it is currently difficult to ascribe these opposite effects to the a putative differential manipulation of A1 and A2A receptors. In animal models aimed at measuring spontaneous anxiety-like responses (such as the light/dark box or the elevated plus maze), there is an anxiogenic-like behaviour in both A1 receptor knockout mice [130,131] as well as in A2A receptor knockout mice [132134]. In contrast, careful studies by our group in CNRS showed that the anxiogenic-like effect of caffeine in rodents is not shared by selective A2A receptor antagonists [135].

Another line of evidence that indicates a possible role of A2A receptors in anxiety-related conditions derives from polymorphism analysis of the A2A receptor gene. Thus, it was observed that there is a significant association between self-reported anxiety after caffeine administration and two linked polymorphisms on the A2A receptor gene, the 1976C>T and 2592C>T polymorphisms [137]. Likewise this same polymorphism in the A2A receptor gene was also observed to be associated with the incidence of panic disorder [137,138], which can be envisage as a situation of anticipatory anxiety. Finally, another polymorphism of the A2A receptor gene (1083TT genotype) is inversely correlated with caffeine consumption [139] and is related with the inter-individual sensitivity to caffeine [140]. This is reminiscent of the idea that there is little evidence for a correlation between the consumption of caffeine and anxiety in volunteers [141,142], but there seems to be an anxiogenic effect of caffeine in a sub-group of patients with different psychiatric disorders [143147]. It remains to be studied if this differential effect of caffeine on anxiety in psychiatric patients may be related to the presence of polymorphisms in the A2A receptor gene [148].

ADENOSINE A2A RECEPTORS AND DEPRESSION

Whether caffeine affects the evolution of depression-like conditions is currently not clear from the epidemiological point of view. In fact, in non-hospitalised cohorts, there is no difference in the consumption of caffeine between control and depressed subjects, albeit there is a trend for greater caffeine-induced anxiety effects in depressed patients [145147]. Likewise, an analysis of life-long caffeine consumption in twin pairs failed to note any evident relation between caffeine intake and the risk for common psychiatric disorders [142].

The association of the adenosine modulation system with depression has been initially developed based on observations showing that adenosine and its analogues caused depressant-like behavioural effects in two widely used animal models of depression. Thus, elevating the adenosine levels increased the time of immobilization in rats submitted to inescapable shocks as well as in the forced swimming test [149151]. Further arguing for an ability of the adenosine system to control depression is the observation that classical antidepressants reverse the adenosine-induced immobility in these tests [152]. Interestingly, classical tricyclic antidepressants such as nortriptiline, chlorimipramine or desipramine can bind to adenosine receptors [153] and dose-dependently reduce the activity of ectonucleotidases in cortical nerve terminals [154], a key controller of the extracellular formation of adenosine from released adenine nucleotides [56]. Accordingly, these tricyclic antidepressants modify the outflow of adenosine from cortical cups [155156] and the glucose and ATP levels in healthy volunteers [157160].

The most direct evidence to implicate adenosine receptors in the control of depression was obtained by our group in CNRS. In a series of careful studies, we found that A2A receptor antagonists prolong escape directed behaviour in two screening tests for antidepressants, the tail suspension and forced swim tests [161]. Further support for a potential role of A2A receptor antagonists as novel anti-depressants was provided by the observation that A2A receptor antagonists also displayed an attenuated ‘behavioural despair’ in these two screening tests [162]. The observation that a dopamine D2 receptor-like antagonist (haloperidol) prevented the antidepressant effects resulting from A2A receptor blockade or inactivation led to the hypothesis these effects of A2A receptors might involve adenosine-dopamine interactions [161,162], in view of the effectiveness of drugs acting on dopaminergic signalling to manage mood disorders. However, additional mechanisms such as the A2A receptor interaction with other neurotransmitter systems in forebrain regions (but outside the striatum) or the ability to control glial metabolism and neuroinflammation should also be explored by future studies.

This putative deleterious role of A2A receptors in depression [162] is in notable agreement with other observations showing that the blockade of A2A receptors relieves the early stress-induced hippocampal modifications [86]. One of the consequences of chronic stress is favouring the implementation of a state of depression in susceptible individuals [163]. Interestingly, adenosine controls the release of corticotrophin and cortisol/corticosterone release [164167] and the ability of adenosine receptor activation to modulate hippocampal excitability [23], a key region in the control of HPA [168], and control memory and cognition, mostly through A2A receptors [20,57,78,169,170]. Finally, adenosine receptors can also control the release of serotonin through A1 and A2A receptors [171] and it has been shown that the ability of caffeine to reduce restraint-induced stress correlates with a striking ability of caffeine to reduce the levels of serotonin in the hippocampus, an effect attributed to A2A receptors [172]. This is particularly relevant since depression as well as the early stress-induced re-modelling of hippocampal circuits are under the control of serotonin (e.g. [173174]) and several novel antidepressant drugs target the serotoninergic system [175].

Another avenue of research that can link A2A receptors with the aetiology of depression resides in the tight interaction between A2A receptors and Trk-B receptors [176], which signal the presence of neurotrophins such as brain-derived neurotrophic factor (BDNF). Thus, there is a continuous build-up and strengthening of the ‘neurotrophin hypothesis of depression’ (reviewed in [177,178]) and evidence is accumulating to suggest that A2A receptors are tight controllers of the actions of BDNF, either through transactivation in an acute manner [179181] or normalization of its signalling in more chronic situations [182]

Furthermore, it is important to keep in mind that the effect of the adenosine modulation system on depressive-like conditions might be more complex. In fact, the group of Ana Lúcia Rodrigues has consistently shown that the administration of adenosine, either peripherally or intracerebroventricularly has an antidepressant effect. This involves the recruitment of A1 and A2A receptors [183] and involves systems such as NO/cGMP [184] or the opioid system [185].

ADENOSINE A2A RECEPTORS AND SCHIZOPHRENIA

Another psychiatric condition where several studies suggest a role for the adenosine modulation system is schizophrenia. Comparing the features of schizophrenia with some physiological roles of adenosine or with the effects of caffeine and theophylline that are used to probe the role of endogenous adenosine, Diogo Lara has championed the idea that adenosinergic activity might be deficient in schizophrenia [186,187]. Thus, caffeine might exacerbate positive symptoms ([188190]; but see [191]) and conversely dipyridamole and allopurinol may be beneficial for schizophrenia [192195]; this provides compelling direct evidence since caffeine blocks adenosine A1 and A2A receptors and both dipyridamole and allopurinol prevent purine degradation by inhibiting adenosine transporters and xanthine oxidase. Furthermore, the expected deficiency of sensorimotor gating, evaluated as a disturbed prepulse inhibition or P50 evoked potential, which is characteristic of schizophrenic individuals [196], is mimicked by theophylline in healthy volunteers [197]. Furthermore, there are co-morbidity relations, namely with insomnia (particularly with delta activity, see [198]), which is mimicked by caffeine consumption [199] and prevented by activation of adenosine receptors [117], and after seizures [200], which is also mimicked by xanthenes and prevented by adenosine A1 receptor activation [201]. Altogether these observations support a putative role for deficient levels of adenosine in the brain of schizophrenic patients and are supportive of the adenosine hypofunction hypothesis of schizophrenia. This hypothesis has been further refined to better match the two-hit hypothesis of schizophrenia, to account for the neuro-developmental aspect of this disorder [186,187]. Thus, A1 receptors have a profound effect of brain development [202], possibly through the control of the function of oligodendrocytes [203206], which would correspond to the first-hit phase. Furthermore, the role of A1 receptors in neuroprotection is only fully implemented during adolescence in rodents [207209], which is compatible with the second hit phase modelling schizophrenia.

In spite of these tempting scenario mainly implying A1 receptors as a candidate system in the aetiology of schizophrenia, there is also compelling observations that suggest a possible role for A2A receptors. Thus, it was observed that the startle (a measure of sensorimotor function) habituation was reduced by A2A receptor antagonists [210] as well as in A2A receptor knockout mice [211]. Furthermore, A2A receptors can also act as ‘go-between’ normalizing (or re-balancing) an impaired glutamatergic-dopaminergic communication that seems to be crucial importance for proper function of the ventral striatum and prefrontal cortex. A recent study with a transgenic model selectively altering the activity of adenosine kinase in forebrain region has provided some direct evidence in supporting the notion that subtle changes in adenosine level can lead to the emergence of behavioural endophenotypes implicated in schizophrenia [212]. Thus, transgenic mice with over-expression of adenosine kinase in the forebrain (to increase adenosine levels) display severe but selective deficits across different learning paradigms, indicating the cognitive function deficient [212]. In addition, altered adenosine level in forebrain also produces abnormal response to psychostimulants, such as amphetamine and MK-801 [212].

Regarding the dopaminergic involvement in schizophrenia, it is noteworthy that activation of adenosine A2A receptors reduces the affinity of dopaminergic D2 receptors for dopamine, being the probable mechanism underlying the antipsychotic-like profile of adenosine agonists [213], the hyperdopaminergic effect of caffeine [100,213] and the exacerbation of psychotic symptoms by caffeine in schizophrenic patients [195]. The recent finding of increased basal D2 receptors occupancy by dopamine in schizophrenic patients [214,215] is compatible with a decreased adenosinergic tone, which via A2A-D2 receptor interaction would increase the affinity of D2 receptors for dopamine [27,213]. Moreover, striatal dopamine release is under tonic inhibition by adenosine acting on presynaptic A1 receptors [216,217], which is also in line with the increased release of dopamine in schizophrenia [218]. Finally, it was observed that the ability of clozapine (an atypical anti-psychotic, an and to a lesser extent haloperidol) to induced c-fos expression is blocked by A2A receptor antagonists [219] and this anti-psychotic also affected key pathways of formation of ATP-derived adenosine acting on A2A receptors, the ecto-nucleotidase pathway [220]. Altogether, these observations are consistent with the possibility that the manipulation of A2A receptor might help restore an adequate dopaminergic signalling.

Concerning the NMDA hypofunction model of schizophrenia [221], adenosine A1 and A2A receptor agonists have been shown to prevent behavioural and EEG effects of NMDA antagonists [222,223]. This effect is in agreement with several lines of evidence: i) activation of NMDA receptors releases adenosine [224228] and ATP [229,230]; ii) administration of NMDA antagonists reduces the basal outflow of adenosine [225,226,228]; iii) the effects of NMDA antagonists may result from increased glutamate release [231233], and both A1 and A2A receptors control the evoked release of glutamate namely in the striatum [35,42,43]; iv) the psychostimulant effects of NMDA receptor antagonists are largely abrogated by genetic or pharmacological blockade of A2A receptors [39,234]; v) NMDA receptor function is modulated both by A1 and by A2A receptors [44,235239]. Taken together, these results suggest that the NMDA hypofunction model may also be corrected by manipulating A2A receptors.

Despite indirect data indicating a potential role for adenosine in the aetiopathology of schizophrenia, direct investigation of the adenosine system in patients is lacking. Acute administration of high doses of caffeine to schizophrenic patients exacerbates positive symptoms but, interestingly, fails to produce anxiety [195,240]. Also, the subtype of adenosine receptor (A1 or A2A) eventually involved in schizophrenia remains undefined. The only post-mortem study of adenosine receptors in schizophrenia reported an increase in striatal A2A receptors [241,242], with no difference between patients on and off medication before death. Also, the A2A receptor gene, located in the 22q12–13 region, is a candidate gene for susceptibility to schizophrenia [243245].

ADENOSINE A2A RECEPTORS AND ADHD

Attention deficit/hyperactivity disorder (ADHD) is a heterogeneous phenotypically complex disorder, whose exact aetiology is unknown. Most probably it does not have a unique cause and represents the final result of different factors that interact with each other, with every factor having a small contribution and increasing the vulnerability to the disorder through their cumulative effects [246,247]. Without underscoring the importance of environmental and psychosocial factors, a substantial genetic component has been detected in the appearance of ADHD, mostly due to data obtained from family, twin and adoption studies [246,248]. Thus, the heritability of ADHD has been estimated to be between 0.5 and 0.9, which makes it the most heritable mental disorder among children. The search for the most probable genetic traits associated with ADHD has mainly targeted genes involved with catecholaminergic transmission, with a special focus on dopamine [249]. Evidence supporting dopaminergic dysfunction in ADHD derives from different research areas: i) first the psychostimulant medication used to counteract ADHD mostly interferes with dopamine transmission [246,250]; ii) behavioural studies in animals indicate a prominent role of dopaminergic transmission in motor control and attention processes [251], which dysfunction are hallmarks of ADHD; iii) neuroimaging studies in ADHD patients demonstrate abnormalities (smaller volumes, hypofunction, decrease blood flow) in brain areas with predominant dopaminergic innervation such as the prefrontal cortex, cingulate gyrus and anterior basal ganglia [252]; iv) case-control and family-based allele frequency studies clearly identified different genes related to dopaminergic transmission (e.g. dopamine receptors and transporter) among the genes associated with higher risk of ADHD [246,248]. In particular, a clear association between ADHD patients and the presence of a particular isoform of the dopamine D4 receptor, the 7R allele (see below), has been extensively replicated (e.g. [253,254]. The fact that this D4–7 receptor allele has a two-fold higher incidence in ADHD probands suggests that it is associated with a significant fraction of the genetic risk for ADHD, which is in accordance with meta-analysis confirming that D4–7 receptor is a susceptibility gene for ADHD [255,256].

This evidence clearly indicates that the D4–7 receptor should be a potential target for the development of novel effective therapeutic strategies to manage ADHD. The D4 receptor belongs to the family of dopamine D2 receptor and displays a number of polymorphisms in Humans, mainly consisting of different repeats in its third exon which encodes the third intracellular loop of D4 receptors; the most common variants have 2, 4 and 7 repeats, which represent more than 90% of the observed allelic diversity [257]. This region is involved in the G protein coupling of D4 receptors and it is interesting to note that the allelic variant that represents a risk factor for ADHD displays a reduced efficacy. Therefore, the therapeutic aim would be to design selective D4 receptor agonists to bolster this defective signalling associated with D4–7 receptor. However, in spite of considerable effort by different research groups, no single compound has yet proven sufficiently potent and selective to activate D4 receptors (e.g. we have found that Ro 10–5824, the most potent and selective D4 receptor agonist available has hitherto unrecognised non- D4 receptor targets in native rodent tissue; unpublished observations). Since D4 receptors belong to the same family as D2 receptors, there is a growing interest in exploring the possibility that A2A receptors may physically interact, not only with D2 receptors (see above), but also with D4 receptors.

The hypothesis that the manipulation of A2A receptors may be a novel therapeutic strategy to manage ADHD is particular compelling in view of the use of caffeine administration to treat this condition [258,259]. In fact, the evidence supporting a dopaminergic dysfunction in ADHD justifies the psychostimulant medication used to counteract ADHD [246,250,260]. Caffeine is the most widely consumed psycho-stimulant drug worldwide and its only known molecular target at non-pathological doses is the antagonism of adenosine receptors, mainly adenosine A1 and A2A receptors [99]. However, the use of caffeine in ADHD is not widespread nor a first choice because it was reported to be less efficient to manage ADHD when compared with other psychostimulant drugs [261]. This contention merits to be revisited in view of the dosage of caffeine used in these studies, which is inadequate to sustain a prolonged blockade of A2A receptors as expected from the pharmacokinetic profile of caffeine [99]. In fact, given that the pharmacokinetic profile of caffeine in children and adolescents indicates a considerably faster elimination of the drug [262265], this once-a-day schedule of caffeine administration is clearly inadequate to provide a plasma level of caffeine sufficient to antagonise A2A receptors throughout the day (in fact, it only allows a 4–6 hours effective antagonism of A2A receptors). Certainly, an adequate use of a novel drug (caffeine), which is innocuous for children [266,267], if effective, would represent a qualitative increment over the traditional repeated use of psychostimulants, which can have severe side effects if repeatedly used in children.

The putative interest of A2A receptors in ADHD has been emphasised by the group of Reinaldo Takahashi, based on the beneficial effects of A2A receptor antagonists in Spontaneous Hypertensive Rats (reviewed in [78]). In fact, it has been shown that these animals have attention deficits that may underlie their poorer memory performance [268270]. Furthermore, these cognitive dysfunctions in SHR are prevented by methylphenidate, which is effective in ADHD [271]. It was observed that caffeine and A2A receptor antagonists are also effective to prevent memory deficits in SHR, while essentially devoid of effects in normal rats [78,272].

CONCLUDING REMARKS

As stressed in the beginning of the review, the lack of clear end–points and of animal models of psychiatric diseases has seriously hampers the ability to critically evaluate the potential of any particular molecule as a relevant target to develop novel drugs to manage psychiatric disorders. The interest in the adenosine system mostly stems from the recognition that its main function is to assist maintaining homeostasis in biological systems. Hence, it should be considered a system of choice to manipulate brain circuits to restore their proper function.

In the particular case of mood disorders, A2A receptors emerge as a promising candidate target since these receptors tightly interact physically and functionally with D2 receptors, which are major targets of psychoactive drugs. The interest on A2A receptors is further emphasised by their prominent role in controlling synaptic plasticity in glutamatergic synapses: thus, a major role of A2A receptors is to normalize the functioning of glutamatergic synapses which dysfunction seems a common feature of many chronic brain diseases. In accordance with this view, A2A receptor blockade affords a robust neuroprotection against different chronic insults to the brain. This neuroprotection afforded by A2A receptor blockade not only depends on the normalization of glutamatergic synapses but also on the ability of A2A receptors to control mitochondria-induced apoptosis as well as to the effectiveness of A2A receptors to control neuro-inflammation. Thus, A2A receptors might not only control the trigger of neuronal dysfunction of brain circuits (glutamate excitotoxicity) but also its main system of amplification (neuroinflammation and metabolic imbalance) as well as its main effector system (apoptotic-induced neuronal damage).

Some caution needs to be introduced in this idyllic scenario. First, there is the need to understand the time window of opportunity to manipulate A2A receptors in brain diseases. There is also an emerging awareness that there are different populations of A2A receptors located in different cellular (and/or sub-cellular) populations that play different and often opposite roles in the control of the function (and dysfunction) of neuronal circuits. In this respect, considerable work still needs to be achieved to allow understanding the molecular mechanisms by which A2A receptors affect brain function. There is growing evidence that A2A receptors are pleiotropic, coupling to different transducing systems, possibly as a function of their heteromerization with different receptors. This opens a thrilling opportunity to manipulate A2A receptors as a novel strategy of “normalisation of signaling” to manage mood disorders.

Finally, there is still an obvious need to validate this potential of A2A receptors where it is in fact relevant, i.e. in patients. This is currently largely restricted to the use of caffeine. Caffeine is known to be a selective adenosine receptor antagonists in rodents (especially in mice), but it might have other hitherto unknown molecular targets in humans. Furthermore, caffeine is not selective for A2A receptors and also antagonises A1 receptors, making it difficult to unambiguously ascribe effects of caffeine as being mediated by A2A receptors. This is hoped to change dramatically in the near future since A2A receptor antagonists have already been approved as novel anti-parkinsonian drugs, which is hoped to bolster our knowledge on the role of A2A receptors in the control of psychiatric disorders.

Acknowledgments

RAC thanks Fundação para a Ciência e Tecnologia and Fundação Oriente for continuous support.Supported also by NIDA IRP funds.

References

  • 1.Belmaker RH, Agam G. Major depressive disorder. N Engl J Med. 2008;358:55–68. doi: 10.1056/NEJMra073096. [DOI] [PubMed] [Google Scholar]
  • 2.Maletic V, Robinson M, Oakes T, Iyengar S, Ball SG, Russell J. Neurobiology of depression: an integrated view of key findings. Int J Clin Pract. 2007;61:2030–40. doi: 10.1111/j.1742-1241.2007.01602.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Nestler EJ, Barrot M, DiLeone RJ, Eisch AJ, Gold SJ, Monteggia LM. Neurobiology of depression. Neuron. 2002;34:13–25. doi: 10.1016/s0896-6273(02)00653-0. [DOI] [PubMed] [Google Scholar]
  • 4.Soares JC, Mann JJ. The functional neuroanatomy of mood disorders. J Psychiatr Res. 1997;31:393–432. doi: 10.1016/s0022-3956(97)00016-2. [DOI] [PubMed] [Google Scholar]
  • 5.Southwick SM, Vythilingam M, Charney DS. The psychobiology of depression and resilience to stress: implications for prevention and treatment. Annu Rev Clin Psychol. 2005;1:255–291. doi: 10.1146/annurev.clinpsy.1.102803.143948. [DOI] [PubMed] [Google Scholar]
  • 6.Fuchs E, Czéh B, Kole MH, Michaelis T, Lucassen PJ. Alterations of neuroplasticity in depression: the hippocampus and beyond. Eur Neuropsychopharmacol. 2004;14(Suppl 5):S481–S490. doi: 10.1016/j.euroneuro.2004.09.002. [DOI] [PubMed] [Google Scholar]
  • 7.Rajkowska G, Miguel-Hidalgo JJ. Gliogenesis and glial pathology in depression. CNS Neurol Disord Drug Targets. 2007;6:219–33. doi: 10.2174/187152707780619326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Hundal Ø. Major depressive disorder viewed as a dysfunction in astroglial bioenergetics. Med Hypotheses. 2007;68:370–377. doi: 10.1016/j.mehy.2006.06.050. [DOI] [PubMed] [Google Scholar]
  • 9.McIntyre RS, Soczynska JK, Konarski JZ, Woldeyohannes HO, Law CW, Miranda A, Fulgosi D, Kennedy SH. Should Depressive Syndromes Be Reclassified as “Metabolic Syndrome Type II”? Ann Clin Psychiatry. 2007;19:257–64. doi: 10.1080/10401230701653377. [DOI] [PubMed] [Google Scholar]
  • 10.Reagan LP. Insulin signaling effects on memory and mood. Curr Opin Pharmacol. 2007;7:633–637. doi: 10.1016/j.coph.2007.10.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Dantzer R. Cytokine, sickness behavior, and depression. Neurol Clin. 2006;24:441–460. doi: 10.1016/j.ncl.2006.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Leonard BE. Inflammation, depression and dementia: are they connected? Neurochem Res. 2007;32:1749–1756. doi: 10.1007/s11064-007-9385-y. [DOI] [PubMed] [Google Scholar]
  • 13.Parent JM. Injury-induced neurogenesis in the adult mammalian brain. Neuroscientist. 2003;9:261–272. doi: 10.1177/1073858403252680. [DOI] [PubMed] [Google Scholar]
  • 14.Steiner B, Wolf S, Kempermann G. Adult neurogenesis and neurodegenerative disease. Regen Med. 2006;1:15–28. doi: 10.2217/17460751.1.1.15. [DOI] [PubMed] [Google Scholar]
  • 15.Cryan JF, Slattery DA. Animal models of mood disorders: Recent developments. Curr Opin Psychiatry. 2007;20:1–7. doi: 10.1097/YCO.0b013e3280117733. [DOI] [PubMed] [Google Scholar]
  • 16.El Yacoubi M, Vaugeois JM. Genetic rodent models of depression. Curr Opin Pharmacol. 2007;7:3–7. doi: 10.1016/j.coph.2006.11.002. [DOI] [PubMed] [Google Scholar]
  • 17.Kalueff AV, Wheaton M, Murphy DL. What’s wrong with my mouse model? Advances and strategies in animal modeling of anxiety and depression. Behav Brain Res. 2007;179:1–18. doi: 10.1016/j.bbr.2007.01.023. [DOI] [PubMed] [Google Scholar]
  • 18.Berton O, Nestler EJ. New approaches to antidepressant drug discovery: beyond monoamines. Nat Rev Neurosci. 2006;7:137–151. doi: 10.1038/nrn1846. [DOI] [PubMed] [Google Scholar]
  • 19.Gould TD, Einat H. Animal models of bipolar disorder and mood stabilizer efficacy: a critical need for improvement. Neurosci Biobehav Rev. 2007;31:825–831. doi: 10.1016/j.neubiorev.2007.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Chen JF, Sonsalla PK, Pedata F, Melani A, Domenici MR, Popoli P, Geiger J, Lopes LV, de Mendonça A. Adenosine A2A receptors and brain injury: broad spectrum of neuroprotection, multifaceted actions and “fine tuning” modulation. Prog Neurobiol. 2007;83:310–31. doi: 10.1016/j.pneurobio.2007.09.002. [DOI] [PubMed] [Google Scholar]
  • 21.Cunha RA. Different cellular sources and different roles of adenosine: A1 receptor-mediated inhibition through astrocytic-driven volume transmission and synapse-restricted A2A receptor-mediated facilitation of plasticity. Neurochem Int. 2008;52:65–72. doi: 10.1016/j.neuint.2007.06.026. [DOI] [PubMed] [Google Scholar]
  • 22.Fredholm BB, Cunha RA, Svenningsson P. Pharmacology of adenosine A2A receptors and therapeutic applications. Curr Top Med Chem. 2003;3:413–426. doi: 10.2174/1568026033392200. [DOI] [PubMed] [Google Scholar]
  • 23.Fredholm BB, Chen JF, Cunha RA, Svenningsson P, Vaugeois JM. Adenosine and brain function. Int Rev Neurobiol. 2005;63:191–270. doi: 10.1016/S0074-7742(05)63007-3. [DOI] [PubMed] [Google Scholar]
  • 24.Ferré S, Fredholm BB, Morelli M, Popoli P, Fuxe K. Adenosine-dopamine receptor-receptor interactions as an integrative mechanism in the basal ganglia. Trends Neurosci. 1997;20:482–487. doi: 10.1016/s0166-2236(97)01096-5. [DOI] [PubMed] [Google Scholar]
  • 25.Moreau JL, Huber G. Central adenosine A2A receptors: an overview. Brain Res Rev. 1999;31:65–82. doi: 10.1016/s0165-0173(99)00059-4. [DOI] [PubMed] [Google Scholar]
  • 26.Richardson PJ, Kase H, Jenner PG. Adenosine A2A receptor antagonists as new agents for the treatment of Parkinson’s disease. Trends Pharmacol Sci. 1997;18:338–344. doi: 10.1016/s0165-6147(97)01096-1. [DOI] [PubMed] [Google Scholar]
  • 27.Svenningsson P, Le Moine C, Fisone G, Fredholm BB. Distribution, biochemistry and function of striatal adenosine A2A receptors. Prog Neurobiol. 1999;59:355–396. doi: 10.1016/s0301-0082(99)00011-8. [DOI] [PubMed] [Google Scholar]
  • 28.Yu L, Shen HY, Coelho JE, Araújo IM, Huang QY, Day YJ, Rebola N, Canas PM, Rapp EK, Ferrara J, Taylor D, Müller CE, Linden J, Cunha RA, Chen JF. A2A receptors modulate motor activity and MPTP neurotoxicity by distinct cellular mechanisms. Ann Neurol. 2008 doi: 10.1002/ana.21313. in press. [DOI] [PubMed] [Google Scholar]
  • 29.Fredholm BB, Chen JF, Masino SA, Vaugeois JM. Actions of adenosine at its receptors in the CNS: insights from knockouts and drugs. Annu Rev Pharmacol Toxicol. 2005;45:385–412. doi: 10.1146/annurev.pharmtox.45.120403.095731. [DOI] [PubMed] [Google Scholar]
  • 30.Schwarzschild MA, Agnati L, Fuxe K, Chen JF, Morelli M. Targeting adenosine A2A receptors in Parkinson’s disease. Trends Neurosci. 2006;29:647–654. doi: 10.1016/j.tins.2006.09.004. [DOI] [PubMed] [Google Scholar]
  • 31.Schiffmann SN, Fisone G, Moresco R, Cunha RA, Ferré S. Adenosine A2A receptors and basal ganglia physiology. Prog Neurobiol. 2007;83:277–292. doi: 10.1016/j.pneurobio.2007.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Gerfen CR. Basal Ganglia. In: Paxinos G, editor. The Rat Nervous System. Elsevier Academic Press; Amsterdam: 2004. pp. 445–508. [Google Scholar]
  • 33.Chen JF, Moratalla R, Impagnatiello F, Grandy DK, Cuellar B, Rubinstein M, Beilstein MA, Hackett E, Fink JS, Low MJ, Ongini E, Schwarzschild MA. The role of the D2 dopamine receptor (D2R) in A2A adenosine receptor (A2AR)-mediated behavioral and cellular responses as revealed by A2A and D2 receptor knockout mice. Proc Natl Acad Sci USA. 2001;98:1970–1975. doi: 10.1073/pnas.98.4.1970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Shiozaki S, Ichikawa S, Nakamura J, Kitamura S, Yamada K, Kuwana Y. Actions of adenosine A2A receptor antagonist KW-6002 on drug-induced catalepsy and hypokinesia caused by reserpine or MPTP. Psychopharmacology. 1999;147:90–95. doi: 10.1007/s002130051146. [DOI] [PubMed] [Google Scholar]
  • 35.Ciruela F, Casadó V, Rodrigues RJ, Luján R, Burgueño J, Canals M, Borycz J, Rebola N, Goldberg SR, Mallol J, Cortés A, Canela EI, López-Giménez JF, Milligan G, Lluis C, Cunha RA, Ferré S, Franco R. Presynaptic control of striatal glutamatergic neurotransmission by adenosine A1-A2A receptor heteromers. J Neurosci. 2006;26:2080–2087. doi: 10.1523/JNEUROSCI.3574-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Bastia E, Xu YH, Scibelli AC, Day YJ, Linden J, Chen JF, Schwarzschild MA. A crucial role for forebrain adenosine A2A receptors in amphetamine sensitization. Neuropsychopharmacology. 2005;30:891–900. doi: 10.1038/sj.npp.1300630. [DOI] [PubMed] [Google Scholar]
  • 37.Huang QY, Wei C, Yu L, Coelho JE, Shen HY, Kalda A, Linden J, Chen JF. Adenosine A2A receptors in bone marrow-derived cells but not in forebrain neurons are important contributors to 3-nitropropionic acid-induced striatal damage as revealed by cell-type-selective inactivation. J Neurosci. 2006;26:11371–11378. doi: 10.1523/JNEUROSCI.1907-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Xiao D, Bastia E, Xu YH, Benn CL, Cha JH, Peterson TS, Chen JF, Schwarzschild MA. Forebrain adenosine A2A receptors contribute to L-3,4-dihydroxyphenylalanine-induced dyskinesia in hemiparkinsonian mice. J Neurosci. 2006;26:13548–13555. doi: 10.1523/JNEUROSCI.3554-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Shen HY, Coelho JE, Ohtsuka N, Canas PM, Day YJ, Huang QY, Rebola N, Yu L, Boison D, Cunha RA, Linden J, Tsien JZ, Chen JF. A critical role of the adenosine A2A receptor in extra-striatal neurons in modulating psychomotor activity as revealed by opposite phenotypes of striatum- and forebrain-A2A receptor knockouts. J Neurosci. 2008 doi: 10.1523/JNEUROSCI.5255-07.2008. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Rebola N, Canas PM, Oliveira CR, Cunha RA. Different synaptic and subsynaptic localization of adenosine A2A receptors in the hippocampus and striatum of the rat. Neuroscience. 2005;132:893–903. doi: 10.1016/j.neuroscience.2005.01.014. [DOI] [PubMed] [Google Scholar]
  • 41.Rebola N, Rodrigues RJ, Lopes LV, Richardson PJ, Oliveira CR, Cunha RA. Adenosine A1 and A2A receptors are co-expressed in pyramidal neurons and co-localized in glutamatergic nerve terminals of the rat hippocampus. Neuroscience. 2005;133:79–83. doi: 10.1016/j.neuroscience.2005.01.054. [DOI] [PubMed] [Google Scholar]
  • 42.Lopes LV, Cunha RA, Kull B, Fredholm BB, Ribeiro JA. Adenosine A2A receptor facilitation of hippocampal synaptic transmission is dependent on tonic A1 receptor inhibition. Neuroscience. 2002;112:319–329. doi: 10.1016/s0306-4522(02)00080-5. [DOI] [PubMed] [Google Scholar]
  • 43.Marchi M, Raiteri L, Risso F, Vallarino A, Bonfanti A, Monopoli A, Ongini E, Raiteri M. Effects of adenosine A1 and A2A receptor activation on the evoked release of glutamate from rat cerebrocortical synaptosomes. Br J Pharmacol. 2002;136:434–440. doi: 10.1038/sj.bjp.0704712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Rebola N, Lujan R, Cunha RA, Mulle C. Adenosine A2A Receptors Are Essential for Long-Term Potentiation of NMDA-EPSCs at Hippocampal Mossy Fiber Synapses. Neuron. 2008;57:121–134. doi: 10.1016/j.neuron.2007.11.023. [DOI] [PubMed] [Google Scholar]
  • 45.Cunha RA, Constantino MD, Ribeiro JA. ZM241385 is an antagonist of the facilitatory responses produced by the A2A adenosine receptor agonists CGS21680 and HENECA in the rat hippocampus. Br J Pharmacol. 1997;122:1279–1284. doi: 10.1038/sj.bjp.0701507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.d’Alcantara P, Ledent C, Swillens S, Schiffmann SN. Inactivation of adenosine A2A receptor impairs long term potentiation in the accumbens nucleus without altering basal synaptic transmission. Neuroscience. 2001;107:455–464. doi: 10.1016/s0306-4522(01)00372-4. [DOI] [PubMed] [Google Scholar]
  • 47.Cunha RA, Correia-de-Sá P, Sebastião AM, Ribeiro JA. Preferential activation of excitatory adenosine receptors at rat hippocampal and neuromuscular synapses by adenosine formed from released adenine nucleotides. Br J Pharmacol. 1996;119:253–260. doi: 10.1111/j.1476-5381.1996.tb15979.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Cunha RA, Vizi ES, Ribeiro JA, Sebastião AM. Preferential release of ATP and its extracellular catabolism as a source of adenosine upon high- but not low-frequency stimulation of rat hippocampal slices. J Neurochem. 1996;67:2180–2187. doi: 10.1046/j.1471-4159.1996.67052180.x. [DOI] [PubMed] [Google Scholar]
  • 49.Cunha RA, Ribeiro JA. Purinergic modulation of [3H]GABA release from rat hippocampal nerve terminals. Neuropharmacology. 2000;39:1156–1167. doi: 10.1016/s0028-3908(99)00237-3. [DOI] [PubMed] [Google Scholar]
  • 50.Kirk IP, Richardson PJ. Inhibition of striatal GABA release by the adenosine A2a receptor is not mediated by increases in cyclic AMP. J Neurochem. 1995;64:2801–2809. doi: 10.1046/j.1471-4159.1995.64062801.x. [DOI] [PubMed] [Google Scholar]
  • 51.Shindou T, Mori A, Kase H, Ichimura M. Adenosine A2A receptor enhances GABA(A)-mediated IPSCs in the rat globus pallidus. J Physiol. 2001;532:423–434. doi: 10.1111/j.1469-7793.2001.0423f.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Shindou T, Richardson PJ, Mori A, Kase H, Ichimura M. Adenosine modulates the striatal GABAergic inputs to the globus pallidus via adenosine A2A receptors in rats. Neurosci Lett. 2003;352:167–170. doi: 10.1016/j.neulet.2003.08.059. [DOI] [PubMed] [Google Scholar]
  • 53.Bacci A, Huguenard JR, Prince DA. Modulation of neocortical interneurons: extrinsic influences and exercises in self-control. Trends Neurosci. 2005;28:602–610. doi: 10.1016/j.tins.2005.08.007. [DOI] [PubMed] [Google Scholar]
  • 54.Cerqueira V, de Mendonça A, Minez A, Dias AR, de Carvalho M. Does caffeine modify corticomotor excitability? Neurophysiol Clin. 2006;36:219–226. doi: 10.1016/j.neucli.2006.08.005. [DOI] [PubMed] [Google Scholar]
  • 55.Cunha RA. Adenosine neuromodulation and neuroprotection. In: Vizi ES, Hamon M, editors. Handbook of Neurochemistry and Molecular Neurobiology. Springer-Verlag; Berlin Heidelberg: 2008. in press. [Google Scholar]
  • 56.Cunha RA. Regulation of the ecto-nucleotidase pathway in rat hippocampal nerve terminals. Neurochem Res. 2001;26:979–991. doi: 10.1023/a:1012392719601. [DOI] [PubMed] [Google Scholar]
  • 57.Cunha RA. Neuroprotection by adenosine in the brain: from A1 receptor activation to A2A receptor blockade. Purinergic Signal. 2005;1:111–134. doi: 10.1007/s11302-005-0649-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Phillis JW. Adenosine and adenine nucleotides as regulators of cerebral blood flow: roles of acidosis, cell swelling, and KATP channels. Crit Rev Neurobiol. 2004;16:237–270. doi: 10.1615/critrevneurobiol.v16.i4.20. [DOI] [PubMed] [Google Scholar]
  • 59.Daré E, Schulte G, Karovic O, Hammarberg C, Fredholm BB. Modulation of glial cell functions by adenosine receptors. Physiol Behav. 2007;92:15–20. doi: 10.1016/j.physbeh.2007.05.031. [DOI] [PubMed] [Google Scholar]
  • 60.Gao Y, Phillis JW. CGS 15943, an adenosine A2 receptor antagonist, reduces cerebral ischemic injury in the Mongolian gerbil. Life Sci. 1994;55:61–65. doi: 10.1016/0024-3205(94)00889-2. [DOI] [PubMed] [Google Scholar]
  • 61.Monopoli A, Lozza G, Forlani A, Mattavelli A, Ongini E. Blockade of adenosine A2A receptors by SCH 58261 results in neuroprotective effects in cerebral ischaemia in rats. Neuroreport. 1998;9:3955–3959. doi: 10.1097/00001756-199812010-00034. [DOI] [PubMed] [Google Scholar]
  • 62.Chen JF, Huang Z, Ma J, Zhu J, Moratalla R, Standaert D, Moskowitz MA, Fink JS, Schwarzschild MA. A(2A) adenosine receptor deficiency attenuates brain injury induced by transient focal ischemia in mice. J Neurosci. 1999;19:9192–200. doi: 10.1523/JNEUROSCI.19-21-09192.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Domenici MR, Scattoni ML, Martire A, Lastoria G, Potenza RL, Borioni A, Venerosi A, Calamandrei G, Popoli P. Behavioral and electrophysiological effects of the adenosine A2A receptor antagonist SCH 58261 in R6/2 Huntington’s disease mice. Neurobiol Dis. 2007;28:197–205. doi: 10.1016/j.nbd.2007.07.009. [DOI] [PubMed] [Google Scholar]
  • 64.Popoli P, Frank C, Tebano MT, Potenza RL, Pintor A, Domenici MR, Nazzicone V, Pèzzola A, Reggio R. Modulation of glutamate release and excitotoxicity by adenosine A2A receptors. Neurology. 2003;61(Suppl 6):S69–S71. doi: 10.1212/01.wnl.0000095216.89483.a2. [DOI] [PubMed] [Google Scholar]
  • 65.Stone TW, Behan WM. Interleukin-1beta but not tumor necrosis factor-alpha potentiates neuronal damage by quinolinic acid: protection by an adenosine A2A receptor antagonist. J Neurosci Res. 2007;85:1077–1085. doi: 10.1002/jnr.21212. [DOI] [PubMed] [Google Scholar]
  • 66.Behan WM, Stone TW. Enhanced neuronal damage by co-administration of quinolinic acid and free radicals, and protection by adenosine A2A receptor antagonists. Br J Pharmacol. 2002;135:1435–1442. doi: 10.1038/sj.bjp.0704613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Jones PA, Smith RA, Stone TW. Protection against hippocampal kainate excitotoxicity by intracerebral administration of an adenosine A2A receptor antagonist. Brain Res. 1998;800:328–335. doi: 10.1016/s0006-8993(98)00540-x. [DOI] [PubMed] [Google Scholar]
  • 68.Lee HK, Choi SS, Han KJ, Han EJ, Suh HW. Roles of adenosine receptors in the regulation of kainic acid-induced neurotoxic responses in mice. Mol Brain Res. 2004;125:76–85. doi: 10.1016/j.molbrainres.2004.03.004. [DOI] [PubMed] [Google Scholar]
  • 69.Zeraati M, Mirnajafi-Zadeh J, Fathollahi Y, Namvar S, Rezvani ME. Adenosine A1 and A2A receptors of hippocampal CA1 region have opposite effects on piriform cortex kindled seizures in rats. Seizure. 2006;15:41–48. doi: 10.1016/j.seizure.2005.10.006. [DOI] [PubMed] [Google Scholar]
  • 70.Chen JF, Xu K, Petzer JP, Staal R, Xu YH, Beilstein M, Sonsalla PK, Castagnoli K, Castagnoli N, Jr, Schwarzschild MA. Neuroprotection by caffeine and A2A adenosine receptor inactivation in a model of Parkinson’s disease. J Neurosci. 2001;21:RC143. doi: 10.1523/JNEUROSCI.21-10-j0001.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Ikeda K, Kurokawa M, Aoyama S, Kuwana Y. Neuroprotection by adenosine A2A receptor blockade in experimental models of Parkinson’s disease. J Neurochem. 2002;80:262–70. doi: 10.1046/j.0022-3042.2001.00694.x. [DOI] [PubMed] [Google Scholar]
  • 72.Xu K, Xu YH, Chen JF, Schwarzschild MA. Caffeine’s neuroprotection against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine toxicity shows no tolerance to chronic caffeine administration in mice. Neurosci Lett. 2002;322:13–16. doi: 10.1016/s0304-3940(02)00069-1. [DOI] [PubMed] [Google Scholar]
  • 73.Blum D, Galas MC, Pintor A, Brouillet E, Ledent C, Muller CE, Bantubungi K, Galluzzo M, Gall D, Cuvelier L, Rolland AS, Popoli P, Schiffmann SN. A dual role of adenosine A2A receptors in 3-nitropropionic acid-induced striatal lesions: implications for the neuroprotective potential of A2A antagonists. J Neurosci. 2003;23:5361–5369. doi: 10.1523/JNEUROSCI.23-12-05361.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Fink JS, Kalda A, Ryu H, Stack EC, Schwarzschild MA, Chen JF, Ferrante RJ. Genetic and pharmacological inactivation of the adenosine A2A receptor attenuates 3-nitropropionic acid-induced striatal damage. J Neurochem. 2004;88:538–544. doi: 10.1046/j.1471-4159.2003.02145.x. [DOI] [PubMed] [Google Scholar]
  • 75.Cunha GMA, Canas PM, Melo CS, Hockemeyer J, Müller CE, Oliveira CR, Cunha RA. Adenosine A2A receptor blockade prevents memory dysfunction caused by β-amyloid peptides but not by scopolamine or MK-801. Exp Neurol. 2008 doi: 10.1016/j.expneurol.2007.11.013. in press. [DOI] [PubMed] [Google Scholar]
  • 76.Dall’Igna OP, Fett P, Gomes MW, Souza DO, Cunha RA, Lara DR. Caffeine and adenosine A2a receptor antagonists prevent beta-amyloid (25–35)-induced cognitive deficits in mice. Exp Neurol. 2007;203:241–245. doi: 10.1016/j.expneurol.2006.08.008. [DOI] [PubMed] [Google Scholar]
  • 77.Lopes LV, Halldner L, Rebola N, Johansson B, Ledent C, Chen JF, Fredholm BB, Cunha RA. Binding of the prototypical adenosine A2A receptor agonist CGS 21680 to the cerebral cortex of adenosine A1 and A2A receptor knockout mice. Br J Pharmacol. 2004;141:1006–1014. doi: 10.1038/sj.bjp.0705692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Takahashi RN, Pamplona FA, Prediger RD. Adenosine receptor antagonists for cognitive dysfunction: a review of animal studies. Front Biosci. 2008;13:2614–2632. doi: 10.2741/2870. [DOI] [PubMed] [Google Scholar]
  • 79.Yu L, Huang Z, Mariani J, Wang Y, Moskowitz M, Chen JF. Selective inactivation or reconstitution of adenosine A2A receptors in bone marrow cells reveals their significant contribution to the development of ischemic brain injury. Nature Med. 2004;10:1081–1087. doi: 10.1038/nm1103. [DOI] [PubMed] [Google Scholar]
  • 80.Cunha RA, Chen JF, Sitkovsky MV. Opposite modulation of peripheral inflammation and neuroinflammation by adenosine A2A receptors. In: Malva JO, Rego AC, Cunha RA, Oliveira CR, editors. Interaction Between Neurons and Glia in Aging and Disease. Springer-Verlag; Berlin: 2007. pp. 53–79. [Google Scholar]
  • 81.Dall’Igna OP, Porciúncula LO, Souza DO, Cunha RA, Lara DR. Neuroprotection by caffeine and adenosine A2A receptor blockade of β-amyloid neurotoxicity. Br J Pharmacol. 2003;138:1207–1209. doi: 10.1038/sj.bjp.0705185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Silva CG, Porciúncula LO, Canas PM, Oliveira CR, Cunha RA. Blockade of adenosine A2A receptors prevents staurosporine-induced apoptosis of rat hippocampal neurons. Neurobiol Dis. 2007;27:182–189. doi: 10.1016/j.nbd.2007.04.018. [DOI] [PubMed] [Google Scholar]
  • 83.Coleman MP, Perry VH. Axon pathology in neurological disease: a neglected therapeutic target. Trends Neurosci. 2002;25:532–537. doi: 10.1016/s0166-2236(02)02255-5. [DOI] [PubMed] [Google Scholar]
  • 84.Selkoe DJ. Alzheimer’s disease is a synaptic failure. Science. 2002;298:789–791. doi: 10.1126/science.1074069. [DOI] [PubMed] [Google Scholar]
  • 85.Wishart TM, Parson SH, Gillingwater TH. Synaptic vulnerability in neurodegenerative disease. J Neuropathol Exp Neurol. 2006;65:733–739. doi: 10.1097/01.jnen.0000228202.35163.c4. [DOI] [PubMed] [Google Scholar]
  • 86.Cunha GM, Canas PM, Oliveira CR, Cunha RA. Increased density and synapto-protective effect of adenosine A2A receptors upon sub-chronic restraint stress. Neuroscience. 2006;141:1775–1781. doi: 10.1016/j.neuroscience.2006.05.024. [DOI] [PubMed] [Google Scholar]
  • 87.Block ML, Hong JS. Microglia and inflammation-mediated neurodegeneration: multiple triggers with a common mechanism. Prog Neurobiol. 2005;76:77–98. doi: 10.1016/j.pneurobio.2005.06.004. [DOI] [PubMed] [Google Scholar]
  • 88.Klegeris A, McGeer EG, McGeer PL. Therapeutic approaches to inflammation in neurodegenerative disease. Curr Opin Neurol. 2007;20:351–357. doi: 10.1097/WCO.0b013e3280adc943. [DOI] [PubMed] [Google Scholar]
  • 89.Rogers J, Mastroeni D, Leonard B, Joyce J, Grover A. Neuroinflammation in Alzheimer’s disease and Parkinson’s disease: are microglia pathogenic in either disorder? Int Rev Neurobiol. 2007;82:235–246. doi: 10.1016/S0074-7742(07)82012-5. [DOI] [PubMed] [Google Scholar]
  • 90.Fredholm BB, Abbracchio MP, Burnstock G, Daly JW, Harden TK, Jacobson KA, Leff P, Williams M. Nomenclature and classification of purinoceptors. Pharmacol Rev. 1994;46:143–156. [PMC free article] [PubMed] [Google Scholar]
  • 91.Hervé D, Le Moine C, Corvol JC, Belluscio L, Ledent C, Fienberg AA, Jaber M, Studler JM, Girault JA. Galpha(olf) levels are regulated by receptor usage and control dopamine and adenosine action in the striatum. J Neurosci. 2001;21:4390–4399. doi: 10.1523/JNEUROSCI.21-12-04390.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kull B, Svenningsson P, Fredholm BB. Adenosine A2A receptors are colocalized with and activate g(olf) in rat striatum. Mol Pharmacol. 2000;58:771–7777. doi: 10.1124/mol.58.4.771. [DOI] [PubMed] [Google Scholar]
  • 93.McPhee I, Gibson LC, Kewney J, Darroch C, Stevens PA, Spinks D, Cooreman A, MacKenzie SJ. Cyclic nucleotide signalling: a molecular approach to drug discovery for Alzheimer’s disease. Biochem Soc Trans. 2005;33:1330–1332. doi: 10.1042/BST0331330. [DOI] [PubMed] [Google Scholar]
  • 94.Tanaka K. Alteration of second messengers during acute cerebral ischemia - adenylate cyclase, cyclic AMP-dependent protein kinase, and cyclic AMP response element binding protein. Prog Neurobiol. 2001;65:173–207. doi: 10.1016/s0301-0082(01)00002-8. [DOI] [PubMed] [Google Scholar]
  • 95.Olesen J, Leonardi M. The burden of brain diseases in Europe. Eur J Neurol. 2003;10:471–477. doi: 10.1046/j.1468-1331.2003.00682.x. [DOI] [PubMed] [Google Scholar]
  • 96.Olesen J, Baker MG, Freund T, di Luca M, Mendlewicz J, Ragan I, Westphal M. Consensus document on European brain research. J Neurol Neurosurg Psychiatry. 2006;77(Suppl 1):i1–49. [PMC free article] [PubMed] [Google Scholar]
  • 97.Sobocki P, Jönsson B, Angst J, Rehnberg C. Cost of depression in Europe. J Ment Health Policy Econ. 2006;9:87–98. [PubMed] [Google Scholar]
  • 98.Smith A. Effects of caffeine on human behavior. Food Chem Toxicol. 2002;40:1243–1255. doi: 10.1016/s0278-6915(02)00096-0. [DOI] [PubMed] [Google Scholar]
  • 99.Fredholm BB, Bättig K, Holmén J, Nehlig A, Zvartau EE. Actions of caffeine in the brain with special reference to factors that contribute to its widespread use. Pharmacol Rev. 1999;51:83–133. [PubMed] [Google Scholar]
  • 100.Ferré S. An update on the mechanisms of the psychostimulant effects of caffeine. J Neurochem. 2008 doi: 10.1111/j.1471-4159.2007.05196.x. in press. [DOI] [PubMed] [Google Scholar]
  • 101.Sawyer DA, Julia HL, Turin AC. Caffeine and human behavior: arousal, anxiety, and performance effects. J Behav Med. 1982;5:415–439. doi: 10.1007/BF00845371. [DOI] [PubMed] [Google Scholar]
  • 102.Gilliland K, Andress D. Ad lib caffeine consumption, symptoms of caffeinism, and academic performance. Am J Psychiatry. 1981;138:512–514. doi: 10.1176/ajp.138.4.512. [DOI] [PubMed] [Google Scholar]
  • 103.Greden JF. Anxiety or caffeinism: a diagnostic dilemma. Am J Psychiatry. 1974;131:1089–1092. doi: 10.1176/ajp.131.10.1089. [DOI] [PubMed] [Google Scholar]
  • 104.Stephenson PE. Physiologic and psychotropic effects of caffeine on man. A review J Am Diet Assoc. 1977;71:240–247. [PubMed] [Google Scholar]
  • 105.Kruger A. Chronic psychiatric patients’ use of caffeine: pharmacological effects and mechanisms. Psychol Rep. 1996;78:915–923. doi: 10.2466/pr0.1996.78.3.915. [DOI] [PubMed] [Google Scholar]
  • 106.Juliano LM, Griffiths RR. A critical review of caffeine withdrawal: empirical validation of symptoms and signs, incidence, severity, and associated features. Psychopharmacology. 2004;176:1–29. doi: 10.1007/s00213-004-2000-x. [DOI] [PubMed] [Google Scholar]
  • 107.Nehlig A, Daval JL, Debry G. Caffeine and the central nervous system: mechanisms of action, biochemical, metabolic and psychostimulant effects. Brain Res Rev. 1992;17:139–170. doi: 10.1016/0165-0173(92)90012-b. [DOI] [PubMed] [Google Scholar]
  • 108.Scher AI, Stewart WF, Lipton RB. Caffeine as a risk factor for chronic daily headache: a population-based study. Neurology. 2004;63:2022–2027. doi: 10.1212/01.wnl.0000145760.37852.ed. [DOI] [PubMed] [Google Scholar]
  • 109.Shapiro RE. Caffeine and headaches. Neurol Sci. 2007;28(Suppl 2):S179–S183. doi: 10.1007/s10072-007-0773-5. [DOI] [PubMed] [Google Scholar]
  • 110.Dews PB, O’Brien CP, Bergman J. Caffeine: behavioral effects of withdrawal and related issues. Food Chem Toxicol. 2002;40:1257–1261. doi: 10.1016/s0278-6915(02)00095-9. [DOI] [PubMed] [Google Scholar]
  • 111.Hughes JR, Oliveto AH, Helzer JE, Higgins ST, Bickel WK. Should caffeine abuse, dependence, or withdrawal be added to DSM-IV and ICD-10? Am J Psychiatry. 1992;149:33–40. doi: 10.1176/ajp.149.1.33. [DOI] [PubMed] [Google Scholar]
  • 112.Strain EC, Mumford GK, Silverman K, Griffiths RR. Caffeine dependence syndrome. Evidence from case histories and experimental evaluations. JAMA. 1994;272:1043–1048. [PubMed] [Google Scholar]
  • 113.van Calker D, Biber K. The role of glial adenosine receptors in neural resilience and the neurobiology of mood disorders. Neurochem Res. 2005;30:1205–1217. doi: 10.1007/s11064-005-8792-1. [DOI] [PubMed] [Google Scholar]
  • 114.Basheer R, Halldner L, Alanko L, McCarley RW, Fredholm BB, Porkka-Heiskanen T. Opposite changes in adenosine A1 and A2A receptor mRNA in the rat following sleep deprivation. Neuroreport. 2001;12:1577–1580. doi: 10.1097/00001756-200106130-00013. [DOI] [PubMed] [Google Scholar]
  • 115.Håberg A, Qu H, Haraldseth O, Unsgård G, Sonnewald U. In vivo effects of adenosine A1 receptor agonist and antagonist on neuronal and astrocytic intermediary metabolism studied with ex vivo 13C NMR spectroscopy. J Neurochem. 2000;74:327–333. doi: 10.1046/j.1471-4159.2000.0740327.x. [DOI] [PubMed] [Google Scholar]
  • 116.Phillis JW, O’Regan MH. Effects of adenosine receptor antagonists on pial arteriolar dilation during carbon dioxide inhalation. Eur J Pharmacol. 2003;476:211–219. doi: 10.1016/s0014-2999(03)02187-3. [DOI] [PubMed] [Google Scholar]
  • 117.Basheer R, Strecker RE, Thakkar MM, McCarley RW. Adenosine and sleep-wake regulation. Prog Neurobiol. 2004;73:379–396. doi: 10.1016/j.pneurobio.2004.06.004. [DOI] [PubMed] [Google Scholar]
  • 118.Gleiter CH, Deckert J, Nutt DJ, Marangos PJ. Electroconvulsive shock (ECS) and the adenosine neuromodulatory system: effect of single and repeated ECS on the adenosine A1 and A2 receptors, adenylate cyclase, and the adenosine uptake site. J Neurochem. 1989;52:641–646. doi: 10.1111/j.1471-4159.1989.tb09168.x. [DOI] [PubMed] [Google Scholar]
  • 119.Yanik G, Radulovacki M. REM sleep deprivation up-regulates adenosine A1 receptors. Brain Res. 1987;402:362–264. doi: 10.1016/0006-8993(87)90046-1. [DOI] [PubMed] [Google Scholar]
  • 120.Kochanek PM, Hendrich KS, Jackson EK, Wisniewski SR, Melick JA, Shore PM, Janesko KL, Zacharia L, Ho C. Characterization of the effects of adenosine receptor agonists on cerebral blood flow in uninjured and traumatically injured rat brain using continuous arterial spin-labeled magnetic resonance imaging. J Cereb Blood Flow Metab. 2005;25:1596–612. doi: 10.1038/sj.jcbfm.9600154. [DOI] [PubMed] [Google Scholar]
  • 121.Ngai AC, Coyne EF, Meno JR, West GA, Winn HR. Receptor subtypes mediating adenosine-induced dilation of cerebral arterioles. Am J Physiol. 2001;280:H2329–H2335. doi: 10.1152/ajpheart.2001.280.5.H2329. [DOI] [PubMed] [Google Scholar]
  • 122.Shin HK, Park SN, Hong KW. Implication of adenosine A2A receptors in hypotension-induced vasodilation and cerebral blood flow autoregulation in rat pial arteries. Life Sci. 2000;67:1435–1445. doi: 10.1016/s0024-3205(00)00737-2. [DOI] [PubMed] [Google Scholar]
  • 123.Green PJ, Suls J. The effects of caffeine on ambulatory blood pressure, heart rate, and mood in coffee drinkers. J Behav Med. 1996;19:111–128. doi: 10.1007/BF01857602. [DOI] [PubMed] [Google Scholar]
  • 124.Lader M, Bruce M. States of anxiety and their induction by drugs. Br J Clin Pharmacol. 1986;22:251–261. doi: 10.1111/j.1365-2125.1986.tb02884.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Loke WH. Effects of caffeine on mood and memory. Physiol Behav. 1988;44:367–372. doi: 10.1016/0031-9384(88)90039-x. [DOI] [PubMed] [Google Scholar]
  • 126.Sicard BA, Perault MC, Enslen M, Chauffard F, Vandel B, Tachon P. The effects of 600 mg of slow release caffeine on mood and alertness. Aviat Space Environ Med. 1996;67:859–862. [PubMed] [Google Scholar]
  • 127.Stern KN, Chait LD, Johanson CE. Reinforcing and subjective effects of caffeine in normal human volunteers. Psychopharmacology. 1989;98:81–88. doi: 10.1007/BF00442010. [DOI] [PubMed] [Google Scholar]
  • 128.Haskell CF, Kennedy DO, Wesnes KA, Scholey AB. Cognitive and mood improvements of caffeine in habitual consumers and habitual non-consumers of caffeine. Psychopharmacology. 2005;179:813–825. doi: 10.1007/s00213-004-2104-3. [DOI] [PubMed] [Google Scholar]
  • 129.Lieberman HR, Tharion WJ, Shukitt-Hale B, Speckman KL, Tulley R. Effects of caffeine, sleep loss, and stress on cognitive performance and mood during U.S. Navy SEAL training. Sea-Air-Land Psychopharmacology. 2002;164:250–261. doi: 10.1007/s00213-002-1217-9. [DOI] [PubMed] [Google Scholar]
  • 130.Giménez-Llort L, Fernández-Teruel A, Escorihuela RM, Fredholm BB, Tobeña A, Pekny M, Johansson B. Mice lacking the adenosine A1 receptor are anxious and aggressive, but are normal learners with reduced muscle strength and survival rate. Eur J Neurosci. 2002;16:547–550. doi: 10.1046/j.1460-9568.2002.02122.x. [DOI] [PubMed] [Google Scholar]
  • 131.Johansson B, Halldner L, Dunwiddie TV, Masino SA, Poelchen W, Giménez-Llort L, Escorihuela RM, Fernández-Teruel A, Wiesenfeld-Hallin Z, Xu XJ, Hårdemark A, Betsholtz C, Herlenius E, Fredholm BB. Hyperalgesia, anxiety, and decreased hypoxic neuroprotection in mice lacking the adenosine A1 receptor. Proc Natl Acad Sci USA. 2001;98:9407–9412. doi: 10.1073/pnas.161292398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Ledent C, Vaugeois JM, Schiffmann SN, Pedrazzini T, El Yacoubi M, Vanderhaeghen JJ, Costentin J, Heath JK, Vassart G, Parmentier M. Aggressiveness, hypoalgesia and high blood pressure in mice lacking the adenosine A2a receptor. Nature. 1997;388:674–678. doi: 10.1038/41771. [DOI] [PubMed] [Google Scholar]
  • 133.Berrendero F, Castañé A, Ledent C, Parmentier M, Maldonado R, Valverde O. Increase of morphine withdrawal in mice lacking A2a receptors and no changes in CB1/A2a double knockout mice. Eur J Neurosci. 2003;17:315–324. doi: 10.1046/j.1460-9568.2003.02439.x. [DOI] [PubMed] [Google Scholar]
  • 134.Bilbao A, Cippitelli A, Martín AB, Granado N, Ortiz O, Bezard E, Chen JF, Navarro M, Rodríguez de Fonseca F, Moratalla R. Absence of quasi-morphine withdrawal syndrome in adenosine A2A receptor knockout mice. Psychopharmacology. 2006;185:160–168. doi: 10.1007/s00213-005-0284-0. [DOI] [PubMed] [Google Scholar]
  • 135.El Yacoubi M, Ledent C, Parmentier M, Costentin J, Vaugeois JM. The anxiogenic-like effect of caffeine in two experimental procedures measuring anxiety in the mouse is not shared by selective A2A adenosine receptor antagonists. Psychopharmacology. 2000;148:153–163. doi: 10.1007/s002130050037. [DOI] [PubMed] [Google Scholar]
  • 136.Alsene K, Deckert J, Sand P, de Wit H. Association between A2a receptor gene polymorphisms and caffeine-induced anxiety. Neuropsychopharmacology. 2003;28:1694–1702. doi: 10.1038/sj.npp.1300232. [DOI] [PubMed] [Google Scholar]
  • 137.Hamilton SP, Slager SL, De Leon AB, Heiman GA, Klein DF, Hodge SE, Weissman MM, Fyer AJ, Knowles JA. Evidence for genetic linkage between a polymorphism in the adenosine 2A receptor and panic disorder. Neuropsychopharmacology. 2004;29:558–565. doi: 10.1038/sj.npp.1300311. [DOI] [PubMed] [Google Scholar]
  • 138.Lam P, Hong CJ, Tsai SJ. Association study of A2a adenosine receptor genetic polymorphism in panic disorder. Neurosci Lett. 2005;378:98–101. doi: 10.1016/j.neulet.2004.12.012. [DOI] [PubMed] [Google Scholar]
  • 139.Cornelis MC, El-Sohemy A, Campos H. Genetic polymorphism of the adenosine A2A receptor is associated with habitual caffeine consumption. Am J Clin Nutr. 2007;86:240–244. doi: 10.1093/ajcn/86.1.240. [DOI] [PubMed] [Google Scholar]
  • 140.Rétey JV, Adam M, Khatami R, Luhmann UF, Jung HH, Berger W, Landolt HP. A genetic variation in the adenosine A2A receptor gene (ADORA2A) contributes to individual sensitivity to caffeine effects on sleep. Clin Pharmacol Ther. 2007;81:692–698. doi: 10.1038/sj.clpt.6100102. [DOI] [PubMed] [Google Scholar]
  • 141.Hire JN. Anxiety and caffeine. Psychol Rep. 1978;42:833–834. doi: 10.2466/pr0.1978.42.3.833. [DOI] [PubMed] [Google Scholar]
  • 142.Kendler KS, Myers JO, Gardner C. Caffeine intake, toxicity and dependence and lifetime risk for psychiatric and substance use disorders: an epidemiologic and co-twin control analysis. Psychol Med. 2006;36:1717–172. doi: 10.1017/S0033291706008622. [DOI] [PubMed] [Google Scholar]
  • 143.Bruce M, Scott N, Shine P, Lader M. Anxiogenic effects of caffeine in patients with anxiety disorders. Arch Gen Psychiatry. 1992;49:867–869. doi: 10.1001/archpsyc.1992.01820110031004. [DOI] [PubMed] [Google Scholar]
  • 144.Charney DS, Heninger GR, Jatlow PI. Increased anxiogenic effects of caffeine in panic disorders. Arch Gen Psychiatry. 1985;42:233–243. doi: 10.1001/archpsyc.1985.01790260027003. [DOI] [PubMed] [Google Scholar]
  • 145.Greden JF, Fontaine P, Lubetsky M, Chamberlin K. Anxiety and depression associated with caffeinism among psychiatric inpatients. Am J Psychiatry. 1987;135:963–966. doi: 10.1176/ajp.135.8.963. [DOI] [PubMed] [Google Scholar]
  • 146.Lee MA, Flegel P, Greden JF, Cameron OG. Anxiogenic effects of caffeine on panic and depressed patients. Am J Psychiatry. 1988;145:632–635. doi: 10.1176/ajp.145.5.632. [DOI] [PubMed] [Google Scholar]
  • 147.Rihs M, Muller C, Baumann P. Caffeine consumption in hospitalized psychiatric patients. Eur Arch Psychiatry Clin Neurosci. 1996;246:83–92. doi: 10.1007/BF02274898. [DOI] [PubMed] [Google Scholar]
  • 148.Tsai SJ, Hong CJ, Hou SJ, Yen FC. Association study of adenosine A2a receptor (1976C>T) genetic polymorphism and mood disorders and age of onset. Psychiatr Genet. 2006;16:185. doi: 10.1097/01.ypg.0000218627.26622.eb. [DOI] [PubMed] [Google Scholar]
  • 149.Hunter AM, Balleine BW, Minor TR. Helplessness and escape performance: glutamate-adenosine interactions in the frontal cortex. Behav Neurosci. 2003;117:123–135. doi: 10.1037//0735-7044.117.1.123. [DOI] [PubMed] [Google Scholar]
  • 150.Minor TR, Winslow JL, Chang WC. Stress and adenosine: II. Adenosine analogs mimic the effect of inescapable shock on shuttle-escape performance in rats. Behav Neurosci. 1994;108:265–276. doi: 10.1037//0735-7044.108.2.265. [DOI] [PubMed] [Google Scholar]
  • 151.Woodson JC, Minor TR, Job RF. Inhibition of adenosine deaminase by erythro-9-(2-hydroxy-3-nonyl)adenine (EHNA) mimics the effect of inescapable shock on escape learning in rats. Behav Neurosci. 1998;112:399–409. doi: 10.1037//0735-7044.112.2.399. [DOI] [PubMed] [Google Scholar]
  • 152.Kulkarni SK, Mehta AK. Purine nucleoside-mediated immobility in mice: reversal by antidepressants. Psychopharmacology. 1985;85:460–463. doi: 10.1007/BF00429665. [DOI] [PubMed] [Google Scholar]
  • 153.Deckert J, Gleiter CH. Adenosinergic psychopharmaceuticals? Trends Pharmacol Sci. 1989;10:99–100. doi: 10.1016/0165-6147(89)90203-4. [DOI] [PubMed] [Google Scholar]
  • 154.Barcellos CK, Schetinger MR, Dias RD, Sarkis JJ. In vitro effect of central nervous system active drugs on the ATPase-ADPase activity and acetylcholinesterase activity from cerebral cortex of adult rats. Gen Pharmacol. 1998;31:563–567. doi: 10.1016/s0306-3623(98)00033-0. [DOI] [PubMed] [Google Scholar]
  • 155.Phillis JW, Wu PH. The effect of various centrally active drugs on adenosine uptake by the central nervous system. Comp Biochem Physiol. 1982;72:179–187. doi: 10.1016/0306-4492(82)90082-x. [DOI] [PubMed] [Google Scholar]
  • 156.Phillis JW. Potentiation of the action of adenosine on cerebral cortical neurons by the tricyclic antidepressants. Br J Pharmacol. 1984;83:567–575. doi: 10.1111/j.1476-5381.1984.tb16521.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Dunn RT, Kimbrell TA, Ketter TA, Frye MA, Willis MW, Luckenbaugh DA, Post RM. Principal components of the Beck Depression Inventory and regional cerebral metabolism in unipolar and bipolar depression. Biol Psychiatry. 2002;51:387–399. doi: 10.1016/s0006-3223(01)01244-6. [DOI] [PubMed] [Google Scholar]
  • 158.Kimbrell TA, Dunn RT, George MS, Danielson AL, Willis MW, Repella JD, Benson BE, Herscovitch P, Post RM, Wassermann EM. Regional cerebral glucose utilization in patients with a range of severities of unipolar depression. Biol Psychiatry. 2002;51:237–252. doi: 10.1016/s0006-3223(01)01216-1. [DOI] [PubMed] [Google Scholar]
  • 159.Renshaw PF, Parow AM, Hirashima F, Ke Y, Moore CM, Frederick Bde B, Fava M, Hennen J, Cohen BM. Multinuclear magnetic resonance spectroscopy studies of brain purines in major depression. Am J Psychiatry. 2001;158:2048–2055. doi: 10.1176/appi.ajp.158.12.2048. [DOI] [PubMed] [Google Scholar]
  • 160.Volz HP, Rzanny R, Riehemann S, May S, Hegewald H, Preussler B, Hübner G, Kaiser WA, Sauer H. 31P magnetic resonance spectroscopy in the frontal lobe of major depressed patients. Eur Arch Psychiatry Clin Neurosci. 1998;248:289–295. doi: 10.1007/s004060050052. [DOI] [PubMed] [Google Scholar]
  • 161.El Yacoubi M, Ledent C, Parmentier M, Bertorelli R, Ongini E, Costentin J, Vaugeois JM. Adenosine A2A receptor antagonists are potential antidepressants: evidence based on pharmacology and A2A receptor knockout mice. Br J Pharmacol. 2001;134:68–77. doi: 10.1038/sj.bjp.0704240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.El Yacoubi M, Costentin J, Vaugeois JM. Adenosine A2A receptors and depression. Neurology. 2003;61(Suppl 6):S82–87. doi: 10.1212/01.wnl.0000095220.87550.f6. [DOI] [PubMed] [Google Scholar]
  • 163.de Kloet ER, Joëls M, Holsboer F. Stress and the brain: from adaptation to disease. Nat Rev Neurosci. 2005;6:463–475. doi: 10.1038/nrn1683. [DOI] [PubMed] [Google Scholar]
  • 164.Chau A, Rose JC, Koos BJ. Adenosine modulates corticotropin and cortisol release during hypoxia in fetal sheep. Am J Obstet Gynecol. 1999;180:1272–1277. doi: 10.1016/s0002-9378(99)70628-9. [DOI] [PubMed] [Google Scholar]
  • 165.Geiger JD, Glavin GB. Adenosine receptor activation in brain reduces stress-induced ulcer formation. Eur J Pharmacol. 1985;115:185–190. doi: 10.1016/0014-2999(85)90690-9. [DOI] [PubMed] [Google Scholar]
  • 166.Jegou S, Yacoubi ME, Mounien L, Ledent C, Parmentier M, Costentin J, Vaugeois JM, Vaudry H. Adenosine A2A receptor gene disruption provokes marked changes in melanocortin content and pro-opiomelanocortin gene expression. J Neuroendocrinol. 2003;15:1171–1177. doi: 10.1111/j.1365-2826.2003.01116.x. [DOI] [PubMed] [Google Scholar]
  • 167.Scaccianoce S, Navarra D, Di Sciullo A, Angelucci L, Endroczi E. Adenosine and pituitary-adrenocortical axis activity in the rat. Neuroendocrinology. 1989;50:464–468. doi: 10.1159/000125264. [DOI] [PubMed] [Google Scholar]
  • 168.Herman JP, Cullinan WE. Neurocircuitry of stress: central control of the hypothalamo-pituitary-adrenocortical axis. Trends Neurosci. 1997;20:78–84. doi: 10.1016/s0166-2236(96)10069-2. [DOI] [PubMed] [Google Scholar]
  • 169.Kopf SR, Melani A, Pedata F, Pepeu G. Adenosine and memory storage: effect of A1 and A2 receptor antagonists. Psychopharmacology. 1999;146:214–219. doi: 10.1007/s002130051109. [DOI] [PubMed] [Google Scholar]
  • 170.Prediger RD, Batista LC, Takahashi RN. Caffeine reverses age-related deficits in olfactory discrimination and social recognition memory in rats. Involvement of adenosine A1 and A2A receptors. Neurobiol Aging. 2005;26:957–964. doi: 10.1016/j.neurobiolaging.2004.08.012. [DOI] [PubMed] [Google Scholar]
  • 171.Okada M, Nutt DJ, Murakami T, Zhu G, Kamata A, Kawata Y, Kaneko S. Adenosine receptor subtypes modulate two major functional pathways for hippocampal serotonin release. J Neurosci. 2001;21:628–640. doi: 10.1523/JNEUROSCI.21-02-00628.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Yamato T, Yamasaki S, Misumi Y, Kino M, Obata T, Aomine M. Modulation of the stress response by coffee: an in vivo microdialysis study of hippocampal serotonin and dopamine levels in rat. Neurosci Lett. 2002;332:87–90. doi: 10.1016/s0304-3940(02)00828-5. [DOI] [PubMed] [Google Scholar]
  • 173.McEwen BS, Conrad CD, Kuroda Y, Frankfurt M, Magarinos AM, McKittrick C. Prevention of stress-induced morphological and cognitive consequences. Eur Neuropsychopharmacol. 1997;7(Suppl 3):S323–S328. doi: 10.1016/s0924-977x(97)00064-3. [DOI] [PubMed] [Google Scholar]
  • 174.McEwen BS, Magarinos AM, Reagan LP. Structural plasticity and tianeptine: cellular and molecular targets. Eur Psychiatry. 2002;17(Suppl 3):318–330. doi: 10.1016/s0924-9338(02)00650-8. [DOI] [PubMed] [Google Scholar]
  • 175.Gillman PK. Tricyclic antidepressant pharmacology and therapeutic drug interactions updated. Br J Pharmacol. 2007;151:737–748. doi: 10.1038/sj.bjp.0707253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Jeanneteau F, Chao MV. Promoting neurotrophic effects by GPCR ligands. Novartis Found Symp. 2006;276:181–189. [PubMed] [Google Scholar]
  • 177.Martinowich K, Manji H, Lu B. New insights into BDNF function in depression and anxiety. Nature Neurosci. 2007;10:1089–1093. doi: 10.1038/nn1971. [DOI] [PubMed] [Google Scholar]
  • 178.Martinowich K, Lu B. Interaction between BDNF and serotonin: role in mood disorders. Neuropsychopharmacology. 2008;33:73–83. doi: 10.1038/sj.npp.1301571. [DOI] [PubMed] [Google Scholar]
  • 179.Diógenes MJ, Fernandes CC, Sebastião AM, Ribeiro JA. Activation of adenosine A2A receptor facilitates brain-derived neurotrophic factor modulation of synaptic transmission in hippocampal slices. J Neurosci. 2004;24:2905–2913. doi: 10.1523/JNEUROSCI.4454-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Lee FS, Chao MV. Activation of Trk neurotrophin receptors in the absence of neurotrophins. Proc Natl Acad Sci USA. 2001;98:3555–3560. doi: 10.1073/pnas.061020198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Tebano MT, Martire A, Potenza RL, Grò C, Pepponi R, Armida M, Domenici MR, Schwarzschild MA, Chen JF, Popoli P. Adenosine A2A receptors are required for normal BDNF levels and BDNF-induced potentiation of synaptic transmission in the mouse hippocampus. J Neurochem. 2008;104:279–286. doi: 10.1111/j.1471-4159.2007.05046.x. [DOI] [PubMed] [Google Scholar]
  • 182.Mojsilovic-Petrovic J, Jeong GB, Crocker A, Arneja A, David S, Russell DS, Kalb RG. Protecting motor neurons from toxic insult by antagonism of adenosine A2a and Trk receptors. J Neurosci. 2006;26:9250–9263. doi: 10.1523/JNEUROSCI.1856-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Kaster MP, Rosa AO, Rosso MM, Goulart EC, Santos AR, Rodrigues AL. Adenosine administration produces an antidepressant-like effect in mice: evidence for the involvement of A1 and A2A receptors. Neurosci Lett. 2004;355:21–24. doi: 10.1016/j.neulet.2003.10.040. [DOI] [PubMed] [Google Scholar]
  • 184.Kaster MP, Rosa AO, Santos AR, Rodrigues AL. Involvement of nitric oxide-cGMP pathway in the antidepressant-like effects of adenosine in the forced swimming test. Int J Neuropsychopharmacol. 2005;8:601–606. doi: 10.1017/S1461145705005316. [DOI] [PubMed] [Google Scholar]
  • 185.Kaster MP, Budni J, Santos AR, Rodrigues AL. Pharmacological evidence for the involvement of the opioid system in the antidepressant-like effect of adenosine in the mouse forced swimming test. Eur J Pharmacol. 2007;576:91–98. doi: 10.1016/j.ejphar.2007.08.026. [DOI] [PubMed] [Google Scholar]
  • 186.Lara DR, Souza DO. Schizophrenia: a purinergic hypothesis. Med Hypotheses. 2000;54:157–166. doi: 10.1054/mehy.1999.0003. [DOI] [PubMed] [Google Scholar]
  • 187.Lara DR, Dall’Igna OP, Ghisolfi ES, Brunstein MG. Involvement of adenosine in the neurobiology of schizophrenia and its therapeutic implications. Prog Neuropsychopharmacol Biol Psychiatry. 2006;30:617–629. doi: 10.1016/j.pnpbp.2006.02.002. [DOI] [PubMed] [Google Scholar]
  • 188.De Freitas B, Schwartz G. Effects of caffeine in chronic psychiatric patients. Am J Psychiatry. 1979;10:1337–1338. doi: 10.1176/ajp.136.10.1337. [DOI] [PubMed] [Google Scholar]
  • 189.Mikkelsen EJ. Caffeine and schizophrenia. J Clin Psychiatry. 1978;9:732–736. [PubMed] [Google Scholar]
  • 190.Nickell PV, Uhde TW. Dose–response effects of intravenous caffeine in normal volunteers. Anxiety. 1994;1:161–168. doi: 10.1002/anxi.3070010403. [DOI] [PubMed] [Google Scholar]
  • 191.Mayo KM, Falkowski W, Jones CA. Caffeine: use and effects in long-stay psychiatric patients. Br J Psychiatry. 1993;162:543–545. doi: 10.1192/bjp.162.4.543. [DOI] [PubMed] [Google Scholar]
  • 192.Akhondzadeh S, Shasavand E, Jamilian H, Shabestari O, Kamalipour A. Dipyridamole in the treatment of schizophrenia: adenosine-dopamine receptor interactions. J Clin Pharm Ther. 2000;2:131–137. doi: 10.1046/j.1365-2710.2000.00273.x. [DOI] [PubMed] [Google Scholar]
  • 193.Akhondzadeh S, Safarcherati A, Amini H. Beneficial antipsychotic effects of allopurinol as add-on therapy for schizophrenia: a double blind, randomized and placebo controlled trial. Prog Neuropsychopharmacol Biol Psychiatry. 2005;2:253–259. doi: 10.1016/j.pnpbp.2004.11.008. [DOI] [PubMed] [Google Scholar]
  • 194.Lara DR, Brunstein MG, Ghisolfi ES, Lobato MI, Belmonte-de-Abreu P, Souza DO. Allopurinol augmentation for poorly responsive schizophrenia. Int Clin Psychopharmacol. 2001;4:235–237. doi: 10.1097/00004850-200107000-00008. [DOI] [PubMed] [Google Scholar]
  • 195.Lucas PB, Pickar D, Kelsoe J, Rapaport M, Pato C, Hommer D. Effects of the acute administration of caffeine in patients with schizophrenia. Biol Psychiatry. 1990;1:35–40. doi: 10.1016/0006-3223(90)90429-6. [DOI] [PubMed] [Google Scholar]
  • 196.Potter D, Summerfelt A, Gold J, Buchanan RW. Review of clinical correlates of P50 sensory gating abnormalities in patients with schizophrenia. Schizophr Bull. 2006;32:692–700. doi: 10.1093/schbul/sbj050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Ghisolfi ES, Prokopiuk AS, Becker J, Ehlers JA, Belmonte-de-Abreu P, Souza DO, Lara DR. The adenosine antagonist theophylline impairs p50 auditory sensory gating in normal subjects. Neuropsychopharmacology. 2002;27:629–637. doi: 10.1016/S0893-133X(02)00337-8. [DOI] [PubMed] [Google Scholar]
  • 198.Keshavan MS, Reynolds CF, III, Miewald MJ, Montrose DM, Sweeney JA, Vasko RC, Jr, Kupfer DJ. Delta sleep deficits in schizophrenia: evidence from automated analyses of sleep data. Arch Gen Psychiatry. 1998;5:443–448. doi: 10.1001/archpsyc.55.5.443. [DOI] [PubMed] [Google Scholar]
  • 199.Landolt HP, Dijk DJ, Gaus SE, Borbely AA. Caffeine reduces low-frequency delta activity in the human sleep EEG. Neuropsychopharmacology. 1995;3:229–238. doi: 10.1016/0893-133X(94)00079-F. [DOI] [PubMed] [Google Scholar]
  • 200.Hyde TM, Weinberger DR. Seizures and schizophrenia. Schizophr Bull. 1997;23:611–622. doi: 10.1093/schbul/23.4.611. [DOI] [PubMed] [Google Scholar]
  • 201.Dunwiddie TV. Adenosine and suppression of seizures. Adv Neurol. 1999;79:1001–1010. [PubMed] [Google Scholar]
  • 202.Rivkees SA, Zhao Z, Porter G, Turner C. Influences of adenosine on the fetus and newborn. Mol Genet Metab. 2001;74:160–171. doi: 10.1006/mgme.2001.3217. [DOI] [PubMed] [Google Scholar]
  • 203.Back SA, Craig A, Luo NL, Ren J, Akundi RS, Ribeiro I, Rivkees SA. Protective effects of caffeine on chronic hypoxia-induced perinatal white matter injury. Ann Neurol. 2006;60:696–705. doi: 10.1002/ana.21008. [DOI] [PubMed] [Google Scholar]
  • 204.Othman T, Yan H, Rivkees SA. Oligodendrocytes express functional A1 adenosine receptors that stimulate cellular migration. Glia. 2003;44:166–172. doi: 10.1002/glia.10281. [DOI] [PubMed] [Google Scholar]
  • 205.Stevens B, Porta S, Haak LL, Gallo V, Fields RD. Adenosine: a neuron-glial transmitter promoting myelination in the CNS in response to action potentials. Neuron. 2002;36:855–868. doi: 10.1016/s0896-6273(02)01067-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Tsutsui S, Schnermann J, Noorbakhsh F, Henry S, Yong VW, Winston BW, Warren K, Power C. A1 adenosine receptor upregulation and activation attenuates neuroinflammation and demyelination in a model of multiple sclerosis. J Neurosci. 2004;24:1521–1529. doi: 10.1523/JNEUROSCI.4271-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Adén U, Herlenius E, Tang LQ, Fredholm BB. Maternal caffeine intake has minor effects on adenosine receptor ontogeny in the rat brain. Pediatr Res. 2000;48:177–183. doi: 10.1203/00006450-200008000-00010. [DOI] [PubMed] [Google Scholar]
  • 208.Dumas TC, Foster TC. Late developmental changes in the ability of adenosine A1 receptors to regulate synaptic transmission in the hippocampus. Dev Brain Res. 1998;105:137–139. [PubMed] [Google Scholar]
  • 209.Rivkees SA. The ontogeny of cardiac and neural A1 adenosine receptor expression in rats. Dev Brain Res. 1995;89:202–213. doi: 10.1016/0165-3806(95)00120-3. [DOI] [PubMed] [Google Scholar]
  • 210.Nagel J, Schladebach H, Koch M, Schwienbacher I, Müller CE, Hauber W. Effects of an adenosine A2A receptor blockade in the nucleus accumbens on locomotion, feeding, and prepulse inhibition in rats. Synapse. 2003;49:279–286. doi: 10.1002/syn.10240. [DOI] [PubMed] [Google Scholar]
  • 211.Wang JH, Short J, Ledent C, Lawrence AJ, Buuse M. Reduced startle habituation and prepulse inhibition in mice lacking the adenosine A2A receptor. Behav Brain Res. 2003;2:201–207. doi: 10.1016/s0166-4328(03)00036-6. [DOI] [PubMed] [Google Scholar]
  • 212.Yee BK, Singer P, Chen JF, Feldon J, Boison D. Transgenic overexpression of adenosine kinase in brain leads to multiple learning impairments and altered sensitivity to psychomimetic drugs. Eur J Neurosci. 2007;26:3237–3252. doi: 10.1111/j.1460-9568.2007.05897.x. [DOI] [PubMed] [Google Scholar]
  • 213.Ferré S. Adenosine–dopamine interactions in the ventral striatum. Implications for the treatment of schizophrenia. Psychopharmacology. 1997;2:107–20. doi: 10.1007/s002130050380. [DOI] [PubMed] [Google Scholar]
  • 214.Abi-Dargham A, Rodenhiser J, Printz D, Zea-Ponce Y, Gil R, Kegeles LS, Weiss R, Cooper TB, Mann JJ, Van Heertum RL, Gorman JM, Laruelle M. Increased baseline occupancy of D2 receptors by dopamine in schizophrenia. Proc Natl Acad Sci USA. 2000;97:8104–8109. doi: 10.1073/pnas.97.14.8104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Seeman P, Schwarz J, Chen JF, Szechtman H, Perreault M, McKnight GS, Roder JC, Quirion R, Boksa P, Srivastava LK, Yanai K, Weinshenker D, Sumiyoshi T. Psychosis pathways converge via D2high dopamine receptors. Synapse. 2006;60:319–346. doi: 10.1002/syn.20303. [DOI] [PubMed] [Google Scholar]
  • 216.Golembiowska K, Zylewska A. Agonists of A1 and A2A adenosine receptors attenuate methamphetamine-induced overflow of dopamine in rat striatum. Brain Res. 1998;2:202–209. doi: 10.1016/s0006-8993(98)00743-4. [DOI] [PubMed] [Google Scholar]
  • 217.Borycz J, Pereira MF, Melani A, Rodrigues RJ, Köfalvi A, Panlilio L, Pedata F, Goldberg SR, Cunha RA, Ferré S. Differential glutamate-dependent and glutamate-independent adenosine A1 receptor-mediated modulation of dopamine release in different striatal compartments. J Neurochem. 2007;101:355–363. doi: 10.1111/j.1471-4159.2006.04386.x. [DOI] [PubMed] [Google Scholar]
  • 218.Laruelle M. The role of endogenous sensitization in the pathophysiology of schizophrenia: implications from recent brain imaging studies. Brain Res Rev. 2000;2–3:371–84. doi: 10.1016/s0165-0173(99)00054-5. [DOI] [PubMed] [Google Scholar]
  • 219.Pinna A, Wardas J, Cozzolino A, Morelli M. Involvement of adenosine A2A receptors in the induction of c-fos expression by clozapine and haloperidol. Neuropsychopharmacology. 1999;20:44–51. doi: 10.1016/S0893-133X(98)00051-7. [DOI] [PubMed] [Google Scholar]
  • 220.Lara DR, Vianna MR, de Paris F, Quevedo J, Oses JP, Battastini AM, Sarkis JJ, Souza DO. Chronic treatment with clozapine, but not haloperidol, increases striatal ecto-5′-nucleotidase activity in rats. Neuropsychobiology. 2001;44:99–102. doi: 10.1159/000054925. [DOI] [PubMed] [Google Scholar]
  • 221.Olney JW, Farber NB. Glutamate receptor dysfunction and schizophrenia. Arch Gen Psychiatry. 1995;12:998–1007. doi: 10.1001/archpsyc.1995.03950240016004. [DOI] [PubMed] [Google Scholar]
  • 222.Popoli P, Reggio R, Pezzola A. Adenosine A1 and A2 receptor agonists significantly prevent the electroencephalographic effects induced by MK-801 in rats. Eur J Pharmacol. 1997;2–3:143–146. doi: 10.1016/s0014-2999(97)01122-9. [DOI] [PubMed] [Google Scholar]
  • 223.Sills TL, Azampanah A, Fletcher PJ. The adenosine A1 receptor agonist N6-cyclopentyladenosine blocks the disruptive effect of phencyclidine on prepulse inhibition of the acoustic startle response in the rat. Eur J Pharmacol. 1999;3:325–9. doi: 10.1016/s0014-2999(99)00088-6. [DOI] [PubMed] [Google Scholar]
  • 224.Chen Y, Graham DI, Stone TW. Release of endogenous adenosine and its metabolites by the activation of NMDA receptors in the rat hippocampus in vivo. Br J Pharmacol. 1992;106:632–638. doi: 10.1111/j.1476-5381.1992.tb14387.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Craig CG, White TD. N-methyl-D-aspartate- and non-N-methyl-D-aspartate-evoked adenosine release from rat cortical slices: distinct purinergic sources and mechanisms of release. J Neurochem. 1993;60:1073–1080. doi: 10.1111/j.1471-4159.1993.tb03256.x. [DOI] [PubMed] [Google Scholar]
  • 226.Manzoni OJ, Manabe T, Nicoll RA. Release of adenosine by activation of NMDA receptors in the hippocampus. Science. 1994;265:2098–2101. doi: 10.1126/science.7916485. [DOI] [PubMed] [Google Scholar]
  • 227.Delaney SM, Geiger JD. Levels of endogenous adenosine in rat striatum. II Regulation of basal and N-methyl-D-aspartate-induced levels by inhibitors of adenosine transport and metabolism. J Pharmacol Exp Ther. 1998;285:568–572. [PubMed] [Google Scholar]
  • 228.Melani A, Corsi C, Giménez-Llort L, Martínez E, Ogren SO, Pedata F, Ferré S. Effect of N-methyl-D-aspartate on motor activity and in vivo adenosine striatal outflow in the rat. Eur J Pharmacol. 1999;385:15–19. doi: 10.1016/s0014-2999(99)00729-3. [DOI] [PubMed] [Google Scholar]
  • 229.Brown P, Dale N. Spike-independent release of ATP from Xenopus spinal neurons evoked by activation of glutamate receptors. J Physiol. 2002;540:851–860. doi: 10.1113/jphysiol.2001.013193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Reigada D, Lu W, Mitchell CH. Glutamate acts at NMDA receptors on fresh bovine and on cultured human retinal pigment epithelial cells to trigger release of ATP. J Physiol. 2006;575:707–720. doi: 10.1113/jphysiol.2006.114439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Moghaddam B, Adams BW. Reversal of phencyclidine effects by a group II metabotropic glutamate receptor agonist in rats. Science. 1998;281:1349–1352. doi: 10.1126/science.281.5381.1349. [DOI] [PubMed] [Google Scholar]
  • 232.Takahata R, Moghaddam B. Activation of glutamate neurotransmission in the prefrontal cortex sustains the motoric and dopaminergic effects of phencyclidine. Neuropsychopharmacology. 2003;28:1117–1124. doi: 10.1038/sj.npp.1300127. [DOI] [PubMed] [Google Scholar]
  • 233.Abekawa T, Ito K, Koyama T. Role of the simultaneous enhancement of NMDA and dopamine D1 receptor-mediated neurotransmission in the effects of clozapine on phencyclidine-induced acute increases in glutamate levels in the rat medial prefrontal cortex. Naunyn Schmiedeberg’s Arch Pharmacol. 2006;374:177–193. doi: 10.1007/s00210-006-0115-9. [DOI] [PubMed] [Google Scholar]
  • 234.Rimondini R, Ferré S, Ogren SO, Fuxe K. Adenosine A2A agonists: a potential new type of atypical antipsychotic. Neuropsychopharmacology. 1997;2:82–91. doi: 10.1016/S0893-133X(97)00033-X. [DOI] [PubMed] [Google Scholar]
  • 235.de Mendonça A, Sebastião AM, Ribeiro JA. Inhibition of NMDA receptor-mediated currents in isolated rat hippocampal neurones by adenosine A1 receptor activation. Neuroreport. 1995;6:1097–1100. doi: 10.1097/00001756-199505300-00006. [DOI] [PubMed] [Google Scholar]
  • 236.Gerevich Z, Wirkner K, Illes P. Adenosine A2A receptors inhibit the N-methyl-D-aspartate component of excitatory synaptic currents in rat striatal neurons. Eur J Pharmacol. 2002;451:161–164. doi: 10.1016/s0014-2999(02)02301-4. [DOI] [PubMed] [Google Scholar]
  • 237.Klishin A, Tsintsadze T, Lozovaya N, Krishtal O. Latent N-methyl-D-aspartate receptors in the recurrent excitatory pathway between hippocampal CA1 pyramidal neurons: Ca2+-dependent activation by blocking A1 adenosine receptors. Proc Natl Acad Sci USA. 1995;92:12431–12435. doi: 10.1073/pnas.92.26.12431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Tebano MT, Martire A, Rebola N, Pepponi R, Domenici MR, Grò MC, Schwarzschild MA, Chen JF, Cunha RA, Popoli P. Adenosine A2A receptors and metabotropic glutamate 5 receptors are co-localized and functionally interact in the hippocampus: a possible key mechanism in the modulation of N-methyl-D-aspartate effects. J Neurochem. 2005;95:1188–1200. doi: 10.1111/j.1471-4159.2005.03455.x. [DOI] [PubMed] [Google Scholar]
  • 239.Wirkner K, Gerevich Z, Krause T, Günther A, Köles L, Schneider D, Nörenberg W, Illes P. Adenosine A2A receptor-induced inhibition of NMDA and GABAA receptor-mediated synaptic currents in a subpopulation of rat striatal neurons. Neuropharmacology. 2004;46:994–1007. doi: 10.1016/j.neuropharm.2004.01.008. [DOI] [PubMed] [Google Scholar]
  • 240.Hughes JR, McHugh P, Holtzman S. Caffeine and schizophrenia. Psychiatr Serv. 1998;11:1415–1417. doi: 10.1176/ps.49.11.1415. [DOI] [PubMed] [Google Scholar]
  • 241.Kurumaji A, Toru M. An increase in [3H] CGS21680 binding in the striatum of postmortem brains of chronic schizophrenics. Brain Res. 1998;808:320–323. doi: 10.1016/s0006-8993(98)00840-3. [DOI] [PubMed] [Google Scholar]
  • 242.Deckert J, Brenner M, Durany N, Zochling R, Paulus W, Ransmayr G, Tatschner T, Danielczyk W, Jellinger K, Riederer P. Up-regulation of striatal adenosine A2A receptors in schizophrenia. NeuroReport. 2003;14:313–316. doi: 10.1097/00001756-200303030-00003. [DOI] [PubMed] [Google Scholar]
  • 243.Deckert J, Nöthen MM, Rietschel M, Wildenauer D, Bondy B, Ertl MA, Knapp M, Schofield PR, Albus M, Maier W, Propping P. Human adenosine A2a receptor (A2aAR) gene: systematic mutation screening in patients with schizophrenia. J Neural Transm. 1996;103:1447–1455. doi: 10.1007/BF01271259. [DOI] [PubMed] [Google Scholar]
  • 244.Hong CJ, Liu HC, Liu TY, Liao DL, Tsai SJ. Association studies of the adenosine A2a receptor (1976T>C) genetic polymorphism in Parkinson’s disease and schizophrenia. J Neural Transm. 2005;112:1503–10. doi: 10.1007/s00702-005-0286-4. [DOI] [PubMed] [Google Scholar]
  • 245.Ottoni GL, Lucchese IC, Martins F, Grillo RW, Bogo MR, Lara DR. Association between 2592C>TINS polymorphism of adenosine A2A receptor gene and schizophrenia. Schizophr Bull. 2005;31:274. [Google Scholar]
  • 246.Burgueño J, Franco R, Ciruela F. Attention Deficit and Hyperactivity disorders. In: Buschmann, Díaz JL, Holenz J, Párraga A, Torrens A, Vela JM, editors. Antidepressants, Antipsychotics, Anxiolytics. Wiley-VCH; 2007. pp. 1090–1182. [Google Scholar]
  • 247.Biederman J, Faraone SV. Attention-deficit hyperactivity disorder. Lancet. 2005;366:237–248. doi: 10.1016/S0140-6736(05)66915-2. [DOI] [PubMed] [Google Scholar]
  • 248.Faraone SV, Perlis RH, Doyle AE, Smoller JW, Goralnick JJ, Holmgren MA, Sklar P. Molecular genetics of attention-deficit/hyperactivity disorder. Biol Psychiatry. 2005;57:1313–1323. doi: 10.1016/j.biopsych.2004.11.024. [DOI] [PubMed] [Google Scholar]
  • 249.Swanson JM, Kinsbourne M, Nigg J, Lanphear B, Stefanatos GA, Volkow N, Taylor E, Casey BJ, Castellanos FX, Wadhwa PD. Etiologic subtypes of attention-deficit/hyperactivity disorder: brain imaging, molecular genetic and environmental factors and the dopamine hypothesis. Neuropsychol Rev. 2007;17:39–59. doi: 10.1007/s11065-007-9019-9. [DOI] [PubMed] [Google Scholar]
  • 250.Wilens TE. Attention deficit hyperactivity disorder and substance use disorders. Am J Psychiatry. 2006;163:2059–2063. doi: 10.1176/ajp.2006.163.12.2059. [DOI] [PubMed] [Google Scholar]
  • 251.Chudasama Y, Robbins TW. Functions of frontostriatal systems in cognition: comparative neuropsychopharmacological studies in rats, monkeys and humans. Biol Psychol. 2006;73:19–38. doi: 10.1016/j.biopsycho.2006.01.005. [DOI] [PubMed] [Google Scholar]
  • 252.Arnsten AF. Fundamentals of attention-deficit/hyperactivity disorder: circuits and pathways. J Clin Psychiatry. 2006;67(Suppl 8):7–12. [PubMed] [Google Scholar]
  • 253.Brookes K, Xu X, Chen W, Zhou K, Neale B, Lowe N, Anney R, Franke B, Gill M, Ebstein R, Buitelaar J, Sham P, Campbell D, Knight J, Andreou P, Altink M, Arnold R, Boer F, Buschgens C, Butler L, Christiansen H, Feldman L, Fleischman K, Fliers E, Howe-Forbes R, Goldfarb A, Heise A, Gabriëls I, Korn-Lubetzki I, Johansson L, Marco R, Medad S, Minderaa R, Mulas F, Müller U, Mulligan A, Rabin K, Rommelse N, Sethna V, Sorohan J, Uebel H, Psychogiou L, Weeks A, Barrett R, Craig I, Banaschewski T, Sonuga-Barke E, Eisenberg J, Kuntsi J, Manor I, McGuffin P, Miranda A, Oades RD, Plomin R, Roeyers H, Rothenberger A, Sergeant J, Steinhausen HC, Taylor E, Thompson M, Faraone SV, Asherson P. The analysis of 51 genes in DSM-IV combined type attention deficit hyperactivity disorder: association signals in DRD4, DAT1 and 16 other genes. Mol Psychiatry. 2006;11:934–953. doi: 10.1038/sj.mp.4001869. [DOI] [PubMed] [Google Scholar]
  • 254.Grady DL, Chi HC, Ding YC, Smith M, Wang E, Schuck S, Flodman P, Spence MA, Swanson JM, Moyzis RK. High prevalence of rare dopamine receptor D4 alleles in children diagnosed with attention-deficit hyperactivity disorder. Mol Psychiatry. 2003;8:536–545. doi: 10.1038/sj.mp.4001350. [DOI] [PubMed] [Google Scholar]
  • 255.Faraone SV, Doyle AE, Mick E, Biederman J. Meta-analysis of the association between the 7-repeat allele of the dopamine D4 receptor gene and attention deficit hyperactivity disorder. Am J Psychiatry. 2001;158:1052–1057. doi: 10.1176/appi.ajp.158.7.1052. [DOI] [PubMed] [Google Scholar]
  • 256.Li D, Sham PC, Owen MJ, He L. Meta-analysis shows significant association between dopamine system genes and attention deficit hyperactivity disorder (ADHD) Hum Mol Genet. 2006;15:2276–2284. doi: 10.1093/hmg/ddl152. [DOI] [PubMed] [Google Scholar]
  • 257.Oak JN, Oldenhof J, Van Tol HH. The dopamine D4 receptor: one decade of research. Eur J Pharmacol. 2000;405:303–327. doi: 10.1016/s0014-2999(00)00562-8. [DOI] [PubMed] [Google Scholar]
  • 258.Gross MD. Caffeine in the treatment of children with minimal brain dysfunction or hyperkinetic syndrome. Psychosomatics. 1975;16:26–27. doi: 10.1016/S0033-3182(75)71230-6. [DOI] [PubMed] [Google Scholar]
  • 259.Garfinkel BD, Webster CD, Sloman L. Responses to methylphenidate and varied doses of caffeine in children with attention deficit disorder. Can J Psychiatry. 1981;26:395–401. doi: 10.1177/070674378102600602. [DOI] [PubMed] [Google Scholar]
  • 260.Arnsten AF. Stimulants: Therapeutic actions in ADHD. Neuropsychopharmacology. 2006;31:2376–2383. doi: 10.1038/sj.npp.1301164. [DOI] [PubMed] [Google Scholar]
  • 261.Huestis RD, Arnold LE, Smeltzer DJ. Caffeine versus methylphenidate and d-amphetamine in minimal brain dysfunction: a double-blind comparison. Am J Psychiatry. 1975;132:868–870. doi: 10.1176/ajp.132.8.868. [DOI] [PubMed] [Google Scholar]
  • 262.Somani SM, Gupta P. Caffeine: a new look at an age-old drug. Int J Clin Pharmacol Ther Toxicol. 1988;26:521–533. [PubMed] [Google Scholar]
  • 263.Pariente-Khayat A, Pons G, Rey E, Richard MO, D’Athis P, Moran C, Badoual J, Olive G. Caffeine acetylator phenotyping during maturation in infants. Pediatr Res. 1991;29:492–495. doi: 10.1203/00006450-199105010-00015. [DOI] [PubMed] [Google Scholar]
  • 264.Lane JR, Connor JD. The influence of endogenous and exogenous sex hormones in adolescents with attention to oral contraceptives and anabolic steroids. J Adolesc Health. 1994;15:630–634. doi: 10.1016/s1054-139x(94)90629-7. [DOI] [PubMed] [Google Scholar]
  • 265.el-Yazigi A, Shabib S, al-Rawithi S, Yusuf A, Legayada ES, al-Humidan A. Salivary clearance and urinary metabolic pattern of caffeine in healthy children and in pediatric patients with hepatocellular diseases. J Clin Pharmacol. 1999;39:366–372. doi: 10.1177/00912709922007930. [DOI] [PubMed] [Google Scholar]
  • 266.Stein MA, Krasowski M, Leventhal BL, Phillips W, Bender BG. Behavioral and cognitive effects of methylxanthines. A meta-analysis of theophylline and caffeine. Arch Pediatr Adolesc Med. 1996;150:284–288. doi: 10.1001/archpedi.1996.02170280054010. [DOI] [PubMed] [Google Scholar]
  • 267.Castellanos FX, Rapoport JL. Effects of caffeine on development and behavior in infancy and childhood: a review of the published literature. Food Chem Toxicol. 2002;40:1235–1242. doi: 10.1016/s0278-6915(02)00097-2. [DOI] [PubMed] [Google Scholar]
  • 268.Alsop B. Problems with spontaneously hypertensive rats (SHR) as a model of attention-deficit/hyperactivity disorder (AD/HD) J Neurosci Methods. 2007;162:42–48. doi: 10.1016/j.jneumeth.2006.12.002. [DOI] [PubMed] [Google Scholar]
  • 269.Jentsch JD. Impaired visuospatial divided attention in the spontaneously hypertensive rat. Behav Brain Res. 2005;157:323–330. doi: 10.1016/j.bbr.2004.07.011. [DOI] [PubMed] [Google Scholar]
  • 270.Li Q, Lu G, Antonio GE, Mak YT, Rudd JA, Fan M, Yew DT. The usefulness of the spontaneously hypertensive rat to model attention-deficit/hyperactivity disorder (ADHD) may be explained by the differential expression of dopamine-related genes in the brain. Neurochem Int. 2007;50:848–857. doi: 10.1016/j.neuint.2007.02.005. [DOI] [PubMed] [Google Scholar]
  • 271.Fox GB, Pan JB, Esbenshade TA, Bennani YL, Black LA, Faghih R, Hancock AA, Decker MW. Effects of histamine H3 receptor ligands GT-2331 and ciproxifan in a repeated acquisition avoidance response in the spontaneously hypertensive rat pup. Behav Brain Res. 2002;131:151–161. doi: 10.1016/s0166-4328(01)00379-5. [DOI] [PubMed] [Google Scholar]
  • 272.Prediger RD, Pamplona FA, Fernandes D, Takahashi RN. Caffeine improves spatial learning deficits in an animal model of attention deficit hyperactivity disorder (ADHD) - the spontaneously hypertensive rat (SHR) Int J Neuropsychopharmacol. 2005;8:583–594. doi: 10.1017/S1461145705005341. [DOI] [PubMed] [Google Scholar]

RESOURCES