ArticlePDF AvailableLiterature Review

Potential Therapeutic Interest of Adenosine A2A Receptors in Psychiatric Disorders

Authors:

Abstract

The interest on targeting adenosine A(2A) receptors in the realm of psychiatric diseases first arose based on their tight physical and functional interaction with dopamine D(2) receptors. However, the role of central A(2A) receptors is now viewed as much broader than just controlling D(2) receptor function. Thus, there is currently a major interest in the ability of A(2A) receptors to control synaptic plasticity at glutamatergic synapses. This is due to a combined ability of A(2A) receptors to facilitate the release of glutamate and the activation of NMDA receptors. Therefore, A(2A) 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 A(2A) receptors particularly attractive targets to manage psychiatric disorders since adenosine may act as go-between glutamate and dopamine, two of the key players in mood processing. Furthermore, A(2A) receptors also control glia function and brain metabolic adaptation, two other emerging mechanisms to understand abnormal processing of mood, and A(2A) receptors are important players 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 A(2A) receptors, based on the effects of caffeine (an antagonist of both A(1) and A(2A) receptors) in psychiatric disorders. However, the introduction of A(2A) receptors antagonists in clinics as anti-parkinsonian agents is hoped to bolster our knowledge on the role of A(2A) receptors in mood disorders in the near future.
Potential therapeutic interest of adenosine A
2A
receptors in
psychiatric disorders
Rodrigo A. Cunha
1
, Sergi Ferré
2
, Jean-Marie Vaugeois
3
, and Jiang-Fan Chen
4
1 Center for Neuroscience of Coimbra, Institute of Biochemistry, Faculty of Medicine, University of Coimbra,
Portugal
2 National Institute on Drug Abuse, I.R.P., N.I.H., D.H.H.S., Baltimore, MD, USA
3 Neuropsychopharmacology of Depression Unit, CNRS FRE 2735 (IFRMP), Institute for Biomedical
Research, University of Rouen, France
4 Department of Neurology, Boston University School of Medicine, USA
Abstract
The interest on targeting adenosine A
2A
receptors in the realm of psychiatric diseases first arose
based on its tight physical and functional interaction with dopamine D
2
receptors. However, the role
of central A
2A
receptors is now viewed as much broader than just controlling D
2
receptor function.
Thus, there is currently a major interest in the ability of A
2A
receptors to control synaptic plasticity
at glutamatergic synapses. This is due to a combined ability of A
2A
receptors to facilitate the release
of glutamate and the activation of NMDA. Therefore, A
2A
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 A
2A
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, A
2A
receptors also control glia function and brain
metabolic adaptation, two other emerging mechanisms to understand abnormal processing of mood,
and A
2A
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 A
2A
receptors, based on the effects of caffeine (an antagonist of both A
1
and
A
2A
receptors) in psychiatric disorders. However, the introduction of A
2A
receptors in clinics as anti-
parkinsonian agents is hoped to bolster our knowledge on the role of A
2A
receptors in mood disorders
in the near future.
Keywords
adenosine; A
2A
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.
NIH Public Access
Author Manuscript
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
Published in final edited form as:
Curr Pharm Des. 2008 ; 14(15): 1512–1524.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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’ [1–5]. 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 [8–
10] 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 [15–17]. 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 A
2A
receptors.
PHYSIOLOGICAL ROLE(S) OF ADENOSINE A
2A
RECEPTORS IN THE BRAIN
There are several reviews dealing with the localization and role of A
2A
receptors in the brain
[20–23]. This short overview is just supposed to recapitulate some features of central A
2A
Cunha et al. Page 2
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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 A
2A
receptors
were confined to the basal ganglia, where they played a role in the control of signal processing
in medium spiny neurons [24–27]. In fact, this particular pool of A
2A
receptors is by far the
most abundant in the mammalian brain, but this should not underscore the fact that A
2A
receptors have a much broader distribution in different brain areas and in different cell types,
albeit with a considerably lower density. These A
2A
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 A
2A
receptor to control motor function [28],
probably the most evident behavioural effect caused by A
2A
receptor ligands [23,29–31].
How these A
2A
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 A
2A
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 A
2A
and dopamine D
2
receptor signalling. However, it is also clear that
A
2A
receptors can control motor function in the absence of dopaminergic signalling [33,34].
This indicates that even striatal A
2A
receptors work through dopamine-independent
mechanisms to impact on brain function. In fact, A
2A
receptors have been localized
presynaptically in a majority of glutamatergic nerve terminals, where they form heteromers
with A
1
receptors and where they play an important facilitatory role of cortico-striatal
glutamatergic neurotransmission [35].
The concept of dopamine-independent effects of A
2A
receptor function is particularly relevant
in the case of extra-striatal A
2A
receptors, where dopaminergic signalling is far less intense.
The most compelling evidence come from the recent study using the brain-region specific
A
2A
receptor knockout models in which A
2A
receptor was selectively deleted either in striatal
neurons (striatum A
2A
KO) or entire forebrain neurons (including striatum, cerebral cortex and
hippocampus, forebrain A
2A
KO) [36–38]. Using these novel knockout models, we recently
showed that cocaine-induced psychomotor activity is enhanced in striatum A
2A
KO mice, but
attenuated in forebrain A
2A
KO mice; urthermore, selective inactivation of A
2A
receptor in
extra-striatal cells by administering the A
2A
receptor antagonist KW6002 to striatum A
2A
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 A
2A
receptors in
extra-striatal neurons in providing a prominent excitatory effect on psychomotor activity
[39]. Theprecise localization of these extra-striatal A
2A
receptors involved in psychomotor is
not clear yet, but several studies have found that these extra-striatal A
2A
receptors are mostly
synaptically-located in contrast to the most abundant striatal A
2A
receptors [40]. In particular,
extra-striatal A
2A
receptors are located in glutamatergic synapses [41]. It is important to point
out that those extra-striatal A
2A
receptors also include the A
2A
receptors localized in striatal
glutamatergic terminals [35] (thus, the term extra-striatal can be a bit misleading). These
A
2A
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
[44–46]. 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 A
2A
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 A
2A
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
Cunha et al. Page 3
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
excitatory circuits through a combined action of A
2A
receptors, as an ancillary system of
NMDA receptors, in synapses engaged in plastic changes, together with the action of inhibitory
A
1
receptors (activated through astrocytic-mediated heterosynaptic depression) in non-
stimulated synapses [21]. Hence A
2A
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 A
2A
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 A
2A
receptors to control inhibitory
transmission in brain circuits. Neurochemical findings showed that A
2A
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 A
2A
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 A
1
and A
2A
receptors to prompt
adaptation and/or restore homeostasis [56,57]. In brain tissues, A
2A
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 A
2A
receptors for brain metabolism still needs to be
thoughtfully tested.
ROLE OF ADENOSINE A
2A
RECEPTORS IN THE CONTROL OF
NEURODEGENERATION
The impact of A
2A
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 A
2A
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 [63–65], free radical
toxicity [66], epilepsy [67–69], MPTP toxicity [70–72], 6-hydroxydopamine toxicity [70,71],
3-nitropropionic acid toxicity [37,73,74] or β-amyloid toxicity [75,76]. Interestingly, the
neuroprotection afforded by A
2A
receptor blockade is most evident in cortical areas (reviewed
in [57]), where the density of A
2A
receptors is nearly 20 times lower than in the striatum
[77]. It is important to note that the neuroprotective effect of A
2A
receptor antagonists in general
correlates with their ability to improve cognitive behaviour in animal models of neurological
disorders [20,57,78]. Consequently, A
2A
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.
Cunha et al. Page 4
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
The mechanism underlying this ability of A
2A
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 A
2A
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 A
2A
receptors from bone marrow-derived cells [37,79], whereas in MPTP-induced neurotoxicity
A
2A
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 A
2A
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 A
2A
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 A
2A
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 A
2A
receptors and with the wide spreading idea that chronic
neurodegenerative diseases begin with synaptic dysfunctions that later evolve into different
demises of neurodegeneration [83–85]. In agreement with this role of synaptic A
2A
receptors
in the control of brain tissue damage is the observation that A
2A
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 A
2A
receptors.
A consensual idea would be to propose that there might be a successive participation of A
2A
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 A
2A
receptors;
with more prolonged noxious stimuli, microglia A
2A
receptors would play a predominant role,
in view of the importance of microglia in the amplification of early brain damage [87–89];
finally, with more severe damage, causing loss of preservation of the blood-brain barrier, it
might be that A
2A
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 A
2A
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 A
2A
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 [91–92]. However, it is now clear
that A
2A
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 A
2A
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 A
2A
receptors (which would trigger but rather prevent accumulations of cAMP)
that actually confers neuroprotection. The clarification of the transducing pathways operated
by A
2A
receptors is an issue of particular relevance since “normalisation of signaling” through
manipulating A
2A
receptors is a potential important issue in the realm of psychiatric disorders.
Cunha et al. Page 5
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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 [95–97]. 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 A
2A
receptors with mood disorders (until data from the use of A
2A
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 A
1
and A
2A
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’ [102–104]. 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 [108–109], withdrawal from
regular consumption of caffeine triggers a variety of anxiety-like symptoms, such as irritability,
sleepiness, dysphoria, nervousness or restlessness [106,107,110–112]. 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 A
1
and facilitatory A
2A
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
Cunha et al. Page 6
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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 A
1
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 A
1
receptors
[117,119] and possibly of A
2A
receptors (reviewed in [57]) the former being a strong candidate
to mediate the reduction of cerebral blood flow [116,120–122], 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 A
2A
RECEPTORS AND ANXIETY
The role of adenosine A
2A
receptors in anxiety is still to be defined. In fact, whereas higher
doses of caffeine tend to increase [103,123–127] 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 A
1
and A
2A
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 A
1
receptor knockout mice [130,131] as
well as in A
2A
receptor knockout mice [132–134]. In contrast, careful studies by our group in
CNRS showed that the anxiogenic-like effect of caffeine in rodents is not shared by selective
A
2A
receptor antagonists [135].
Another line of evidence that indicates a possible role of A
2A
receptors in anxiety-related
conditions derives from polymorphism analysis of the A
2A
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 A
2A
receptor gene, the 1976C>T and 2592C>T
polymorphisms [137]. Likewise this same polymorphism in the A
2A
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 A
2A
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 [143–147]. 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 A
2A
receptor gene [148].
ADENOSINE A
2A
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 [145–147]. 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 [149–151]. Further arguing for an ability of the adenosine
system to control depression is the observation that classical antidepressants reverse the
Cunha et al. Page 7
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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 [155–156] and the glucose and ATP levels in healthy volunteers
[157–160].
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 A
2A
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 A
2A
receptor
antagonists as novel anti-depressants was provided by the observation that A
2A
receptor
antagonists also displayed an attenuated ‘behavioural despair’ in these two screening tests
[162]. The observation that a dopamine D
2
receptor-like antagonist (haloperidol) prevented
the antidepressant effects resulting from A
2A
receptor blockade or inactivation led to the
hypothesis these effects of A
2A
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 A
2A
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 A
2A
receptors in depression [162] is in notable agreement with
other observations showing that the blockade of A
2A
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 [164–167]
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
A
2A
receptors [20,57,78,169,170]. Finally, adenosine receptors can also control the release of
serotonin through A
1
and A
2A
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 A
2A
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. [173–174]) and several novel
antidepressant drugs target the serotoninergic system [175].
Another avenue of research that can link A
2A
receptors with the aetiology of depression resides
in the tight interaction between A
2A
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 A
2A
receptors
are tight controllers of the actions of BDNF, either through transactivation in an acute manner
[179–181] 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 A
1
and
A
2A
receptors [183] and involves systems such as NO/cGMP [184] or the opioid system
[185].
Cunha et al. Page 8
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
ADENOSINE A
2A
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 ([188–190]; but see [191]) and conversely dipyridamole and
allopurinol may be beneficial for schizophrenia [192–195]; this provides compelling direct
evidence since caffeine blocks adenosine A
1
and A
2A
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 A
1
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, A
1
receptors have a profound effect of
brain development [202], possibly through the control of the function of oligodendrocytes
[203–206], which would correspond to the first-hit phase. Furthermore, the role of A
1
receptors
in neuroprotection is only fully implemented during adolescence in rodents [207–209], which
is compatible with the second hit phase modelling schizophrenia.
In spite of these tempting scenario mainly implying A
1
receptors as a candidate system in the
aetiology of schizophrenia, there is also compelling observations that suggest a possible role
for A
2A
receptors. Thus, it was observed that the startle (a measure of sensorimotor function)
habituation was reduced by A
2A
receptor antagonists [210] as well as in A
2A
receptor knockout
mice [211]. Furthermore, A
2A
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 A
2A
receptors reduces the affinity of dopaminergic D
2
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
D
2
receptors occupancy by dopamine in schizophrenic patients [214,215] is compatible with
a decreased adenosinergic tone, which via A
2A
-D
2
receptor interaction would increase the
affinity of D
2
receptors for dopamine [27,213]. Moreover, striatal dopamine release is under
tonic inhibition by adenosine acting on presynaptic A
1
receptors [216,217], which is also in
line with the increased release of dopamine in schizophrenia [218]. Finally, it was observed
Cunha et al. Page 9
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
that the ability of clozapine (an atypical anti-psychotic, an and to a lesser extent haloperidol)
to induced c-fos expression is blocked by A
2A
receptor antagonists [219] and this anti-psychotic
also affected key pathways of formation of ATP-derived adenosine acting on A
2A
receptors,
the ecto-nucleotidase pathway [220]. Altogether, these observations are consistent with the
possibility that the manipulation of A
2A
receptor might help restore an adequate dopaminergic
signalling.
Concerning the NMDA hypofunction model of schizophrenia [221], adenosine A
1
and A
2A
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 [224–228] 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 [231–233], and both A
1
and
A
2A
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 A
2A
receptors [39,234]; v) NMDA receptor function is
modulated both by A
1
and by A
2A
receptors [44,235–239]. Taken together, these results
suggest that the NMDA hypofunction model may also be corrected by manipulating A
2A
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 (A
1
or A
2A
) eventually involved in schizophrenia remains undefined. The only post-
mortem study of adenosine receptors in schizophrenia reported an increase in striatal A
2A
receptors [241,242], with no difference between patients on and off medication before death.
Also, the A
2A
receptor gene, located in the 22q12–13 region, is a candidate gene for
susceptibility to schizophrenia [243–245].
ADENOSINE A
2A
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
Cunha et al. Page 10
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
ADHD patients and the presence of a particular isoform of the dopamine D
4
receptor, the 7R
allele (see below), has been extensively replicated (e.g. [253,254]. The fact that this D
4–7
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 D
4–7
receptor is a susceptibility gene for ADHD [255,256].
This evidence clearly indicates that the D
4–7
receptor should be a potential target for the
development of novel effective therapeutic strategies to manage ADHD. The D
4
receptor
belongs to the family of dopamine D
2
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 D
4
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 D
4
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 D
4
receptor agonists to bolster this defective signalling associated
with D
4–7
receptor. However, in spite of considerable effort by different research groups, no
single compound has yet proven sufficiently potent and selective to activate D
4
receptors (e.g.
we have found that Ro 10–5824, the most potent and selective D
4
receptor agonist available
has hitherto unrecognised non- D
4
receptor targets in native rodent tissue; unpublished
observations). Since D
4
receptors belong to the same family as D
2
receptors, there is a growing
interest in exploring the possibility that A
2A
receptors may physically interact, not only with
D
2
receptors (see above), but also with D
4
receptors.
The hypothesis that the manipulation of A
2A
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
A
1
and A
2A
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 A
2A
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 [262–265], this once-a-day schedule of caffeine administration is
clearly inadequate to provide a plasma level of caffeine sufficient to antagonise A
2A
receptors
throughout the day (in fact, it only allows a 4–6 hours effective antagonism of A
2A
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 A
2A
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 [268–270]. Furthermore,
these cognitive dysfunctions in SHR are prevented by methylphenidate, which is effective in
ADHD [271]. It was observed that caffeine and A
2A
receptor antagonists are also effective to
prevent memory deficits in SHR, while essentially devoid of effects in normal rats [78,272].
Cunha et al. Page 11
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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, A
2A
receptors emerge as a promising candidate target
since these receptors tightly interact physically and functionally with D
2
receptors, which are
major targets of psychoactive drugs. The interest on A
2A
receptors is further emphasised by
their prominent role in controlling synaptic plasticity in glutamatergic synapses: thus, a major
role of A
2A
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, A
2A
receptor blockade affords a robust neuroprotection against different chronic insults
to the brain. This neuroprotection afforded by A
2A
receptor blockade not only depends on the
normalization of glutamatergic synapses but also on the ability of A
2A
receptors to control
mitochondria-induced apoptosis as well as to the effectiveness of A
2A
receptors to control
neuro-inflammation. Thus, A
2A
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 A
2A
receptors in brain diseases.
There is also an emerging awareness that there are different populations of A
2A
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 A
2A
receptors affect brain function. There is growing evidence that
A
2A
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 A
2A
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 A
2A
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 A
2A
receptors and also antagonises A
1
receptors, making it difficult to
unambiguously ascribe effects of caffeine as being mediated by A
2A
receptors. This is hoped
to change dramatically in the near future since A
2A
receptor antagonists have already been
approved as novel anti-parkinsonian drugs, which is hoped to bolster our knowledge on the
role of A
2A
receptors in the control of psychiatric disorders.
Acknowledgements
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. [PubMed:
18172175]
Cunha et al. Page 12
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. [PubMed: 17944926]
3. Nestler EJ, Barrot M, DiLeone RJ, Eisch AJ, Gold SJ, Monteggia LM. Neurobiology of depression.
Neuron 2002;34:13–25. [PubMed: 11931738]
4. Soares JC, Mann JJ. The functional neuroanatomy of mood disorders. J Psychiatr Res 1997;31:393–
432. [PubMed: 9352470]
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. [PubMed:
17716089]
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. [PubMed:
15550346]
7. Rajkowska G, Miguel-Hidalgo JJ. Gliogenesis and glial pathology in depression. CNS Neurol Disord
Drug Targets 2007;6:219–33. [PubMed: 17511618]
8. Hundal Ø. Major depressive disorder viewed as a dysfunction in astroglial bioenergetics. Med
Hypotheses 2007;68:370–377. [PubMed: 16978794]
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. [PubMed: 18058283]
10. Reagan LP. Insulin signaling effects on memory and mood. Curr Opin Pharmacol 2007;7:633–637.
[PubMed: 18023616]
11. Dantzer R. Cytokine, sickness behavior, and depression. Neurol Clin 2006;24:441–460. [PubMed:
16877117]
12. Leonard BE. Inflammation, depression and dementia: are they connected? Neurochem Res
2007;32:1749–1756. [PubMed: 17705097]
13. Parent JM. Injury-induced neurogenesis in the adult mammalian brain. Neuroscientist 2003;9:261–
272. [PubMed: 12934709]
14. Steiner B, Wolf S, Kempermann G. Adult neurogenesis and neurodegenerative disease. Regen Med
2006;1:15–28. [PubMed: 17465817]
15. Cryan JF, Slattery DA. Animal models of mood disorders: Recent developments. Curr Opin
Psychiatry 2007;20:1–7. [PubMed: 17143074]
16. El Yacoubi M, Vaugeois JM. Genetic rodent models of depression. Curr Opin Pharmacol 2007;7:3–
7. [PubMed: 17169613]
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. [PubMed:
17306892]
18. Berton O, Nestler EJ. New approaches to antidepressant drug discovery: beyond monoamines. Nat
Rev Neurosci 2006;7:137–151. [PubMed: 16429123]
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. [PubMed: 17628675]
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. [PubMed: 18023959]
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. [PubMed: 17664029]
22. Fredholm BB, Cunha RA, Svenningsson P. Pharmacology of adenosine A2A receptors and
therapeutic applications. Curr Top Med Chem 2003;3:413–426. [PubMed: 12570759]
23. Fredholm BB, Chen JF, Cunha RA, Svenningsson P, Vaugeois JM. Adenosine and brain function.
Int Rev Neurobiol 2005;63:191–270. [PubMed: 15797469]
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.
[PubMed: 9347617]
Cunha et al. Page 13
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
25. Moreau JL, Huber G. Central adenosine A2A receptors: an overview. Brain Res Rev 1999;31:65–
82. [PubMed: 10611496]
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. [PubMed: 9345853]
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. [PubMed: 10501634]
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. 2008in press
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. [PubMed:
15822182]
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. [PubMed: 17030429]
31. Schiffmann SN, Fisone G, Moresco R, Cunha RA, Ferré S. Adenosine A2A receptors and basal
ganglia physiology. Prog Neurobiol 2007;83:277–292. [PubMed: 17646043]
32. Gerfen, CR. Basal Ganglia. In: Paxinos, G., editor. The Rat Nervous System. Elsevier Academic
Press; Amsterdam: 2004. p. 445-508.
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. [PubMed: 11172060]
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. [PubMed: 10591873]
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. [PubMed: 16481441]
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. [PubMed: 15602504]
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. [PubMed: 17079665]
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. [PubMed: 17192438]
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. 2008in press
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. [PubMed: 15857695]
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. [PubMed: 15893632]
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. [PubMed: 12044450]
Cunha et al. Page 14
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. [PubMed: 12023946]
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.
[PubMed: 18184569]
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. [PubMed: 9421273]
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. [PubMed: 11719000]
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. [PubMed: 8886406]
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. [PubMed: 8863529]
49. Cunha RA, Ribeiro JA. Purinergic modulation of [3H]GABA release from rat hippocampal nerve
terminals. Neuropharmacology 2000;39:1156–1167. [PubMed: 10760359]
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. [PubMed: 7760061]
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. [PubMed: 11306661]
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.
[PubMed: 14625011]
53. Bacci A, Huguenard JR, Prince DA. Modulation of neocortical interneurons: extrinsic influences and
exercises in self-control. Trends Neurosci 2005;28:602–610. [PubMed: 16139371]
54. Cerqueira V, de Mendonça A, Minez A, Dias AR, de Carvalho M. Does caffeine modify corticomotor
excitability? Neurophysiol Clin 2006;36:219–226. [PubMed: 17095411]
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
56. Cunha RA. Regulation of the ecto-nucleotidase pathway in rat hippocampal nerve terminals.
Neurochem Res 2001;26:979–991. [PubMed: 11699950]
57. Cunha RA. Neuroprotection by adenosine in the brain: from A1 receptor activation to A2A receptor
blockade. Purinergic Signal 2005;1:111–134. [PubMed: 18404497]
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. [PubMed: 15862108]
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. [PubMed: 17574632]
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.
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. [PubMed: 9875735]
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. [PubMed: 10531422]
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. [PubMed:
17720507]
Cunha et al. Page 15
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. [PubMed: 14663014]
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. [PubMed: 17304576]
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. [PubMed: 11906956]
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.
[PubMed: 9685693]
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. [PubMed: 15193424]
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. [PubMed: 16337818]
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. [PubMed: 11319241]
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. [PubMed: 11902116]
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. [PubMed: 11958832]
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. [PubMed: 12832562]
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. [PubMed: 14720203]
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. 2008in press
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. [PubMed: 17007839]
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. [PubMed: 14993095]
78. Takahashi RN, Pamplona FA, Prediger RD. Adenosine receptor antagonists for cognitive dysfunction:
a review of animal studies. Front Biosci 2008;13:2614–2632. [PubMed: 17981738]
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. [PubMed: 15448683]
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. p. 53-79.
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.
[PubMed: 12711619]
Cunha et al. Page 16
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. [PubMed: 17596953]
83. Coleman MP, Perry VH. Axon pathology in neurological disease: a neglected therapeutic target.
Trends Neurosci 2002;25:532–537. [PubMed: 12220882]
84. Selkoe DJ. Alzheimer’s disease is a synaptic failure. Science 2002;298:789–791. [PubMed:
12399581]
85. Wishart TM, Parson SH, Gillingwater TH. Synaptic vulnerability in neurodegenerative disease. J
Neuropathol Exp Neurol 2006;65:733–739. [PubMed: 16896307]
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.
[PubMed: 16797134]
87. Block ML, Hong JS. Microglia and inflammation-mediated neurodegeneration: multiple triggers with
a common mechanism. Prog Neurobiol 2005;76:77–98. [PubMed: 16081203]
88. Klegeris A, McGeer EG, McGeer PL. Therapeutic approaches to inflammation in neurodegenerative
disease. Curr Opin Neurol 2007;20:351–357. [PubMed: 17495632]
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. [PubMed: 17678964]
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. [PubMed:
7938164]
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. [PubMed: 11404425]
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. [PubMed: 10999947]
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. [PubMed: 16246111]
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. [PubMed: 11403878]
95. Olesen J, Leonardi M. The burden of brain diseases in Europe. Eur J Neurol 2003;10:471–477.
[PubMed: 12940825]
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. [PubMed:
16845120]
97. Sobocki P, Jönsson B, Angst J, Rehnberg C. Cost of depression in Europe. J Ment Health Policy Econ
2006;9:87–98. [PubMed: 17007486]
98. Smith A. Effects of caffeine on human behavior. Food Chem Toxicol 2002;40:1243–1255. [PubMed:
12204388]
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:
10049999]
100. Ferré S. An update on the mechanisms of the psychostimulant effects of caffeine. J Neurochem.
2008in press
101. Sawyer DA, Julia HL, Turin AC. Caffeine and human behavior: arousal, anxiety, and performance
effects. J Behav Med 1982;5:415–439. [PubMed: 7154064]
102. Gilliland K, Andress D. Ad lib caffeine consumption, symptoms of caffeinism, and academic
performance. Am J Psychiatry 1981;138:512–514. [PubMed: 7212112]
103. Greden JF. Anxiety or caffeinism: a diagnostic dilemma. Am J Psychiatry 1974;131:1089–1092.
[PubMed: 4412061]
Cunha et al. Page 17
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
104. Stephenson PE. Physiologic and psychotropic effects of caffeine on man. A review J Am Diet Assoc
1977;71:240–247.
105. Kruger A. Chronic psychiatric patients’ use of caffeine: pharmacological effects and mechanisms.
Psychol Rep 1996;78:915–923. [PubMed: 8711047]
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.
[PubMed: 15448977]
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. [PubMed:
1356551]
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. [PubMed: 15596744]
109. Shapiro RE. Caffeine and headaches. Neurol Sci 2007;28(Suppl 2):S179–S183. [PubMed:
17508167]
110. Dews PB, O’Brien CP, Bergman J. Caffeine: behavioral effects of withdrawal and related issues.
Food Chem Toxicol 2002;40:1257–1261. [PubMed: 12204389]
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. [PubMed:
1728182]
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:
8089887]
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. [PubMed: 16341582]
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. [PubMed: 11409719]
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. [PubMed: 10617136]
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. [PubMed: 12969768]
117. Basheer R, Strecker RE, Thakkar MM, McCarley RW. Adenosine and sleep-wake regulation. Prog
Neurobiol 2004;73:379–396. [PubMed: 15313333]
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. [PubMed:
2911034]
119. Yanik G, Radulovacki M. REM sleep deprivation up-regulates adenosine A1 receptors. Brain Res
1987;402:362–264. [PubMed: 3030496]
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. [PubMed: 15931163]
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.
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.
[PubMed: 10983840]
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. [PubMed: 9132505]
124. Lader M, Bruce M. States of anxiety and their induction by drugs. Br J Clin Pharmacol 1986;22:251–
261. [PubMed: 3533122]
Cunha et al. Page 18
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
125. Loke WH. Effects of caffeine on mood and memory. Physiol Behav 1988;44:367–372. [PubMed:
3222359]
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:
9025802]
127. Stern KN, Chait LD, Johanson CE. Reinforcing and subjective effects of caffeine in normal human
volunteers. Psychopharmacology 1989;98:81–88. [PubMed: 2498963]
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. [PubMed: 15678363]
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.
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.
[PubMed: 12193199]
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. [PubMed: 11470917]
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. [PubMed: 9262401]
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. [PubMed: 12542668]
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. [PubMed: 16470403]
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. [PubMed:
10663430]
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. [PubMed: 12825092]
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. [PubMed: 14666117]
138. Lam P, Hong CJ, Tsai SJ. Association study of A2a adenosine receptor genetic polymorphism in
panic disorder. Neurosci Lett 2005;378:98–101. [PubMed: 15774265]
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. [PubMed:
17616786]
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. [PubMed: 17329997]
141. Hire JN. Anxiety and caffeine. Psychol Rep 1978;42:833–834. [PubMed: 674508]
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. [PubMed: 16893482]
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. [PubMed: 1444724]
Cunha et al. Page 19
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
144. Charney DS, Heninger GR, Jatlow PI. Increased anxiogenic effects of caffeine in panic disorders.
Arch Gen Psychiatry 1985;42:233–243. [PubMed: 2983630]
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. [PubMed: 665843]
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. [PubMed: 3358468]
147. Rihs M, Muller C, Baumann P. Caffeine consumption in hospitalized psychiatric patients. Eur Arch
Psychiatry Clin Neurosci 1996;246:83–92. [PubMed: 9063913]
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. [PubMed:
16969271]
149. Hunter AM, Balleine BW, Minor TR. Helplessness and escape performance: glutamate-adenosine
interactions in the frontal cortex. Behav Neurosci 2003;117:123–135. [PubMed: 12619915]
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.
[PubMed: 8037870]
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. [PubMed: 9588486]
152. Kulkarni SK, Mehta AK. Purine nucleoside-mediated immobility in mice: reversal by
antidepressants. Psychopharmacology 1985;85:460–463. [PubMed: 2991960]
153. Deckert J, Gleiter CH. Adenosinergic psychopharmaceuticals? Trends Pharmacol Sci 1989;10:99–
100. [PubMed: 2595795]
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. [PubMed: 9792215]
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.
156. Phillis JW. Potentiation of the action of adenosine on cerebral cortical neurons by the tricyclic
antidepressants. Br J Pharmacol 1984;83:567–575. [PubMed: 6487906]
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. [PubMed: 11904133]
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. [PubMed: 11839367]
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. [PubMed: 11729024]
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. [PubMed: 9928907]
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. [PubMed: 11522598]
162. El Yacoubi M, Costentin J, Vaugeois JM. Adenosine A2A receptors and depression. Neurology
2003;61(Suppl 6):S82–87. [PubMed: 14663017]
163. de Kloet ER, Joëls M, Holsboer F. Stress and the brain: from adaptation to disease. Nat Rev Neurosci
2005;6:463–475. [PubMed: 15891777]
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. [PubMed: 10329889]
165. Geiger JD, Glavin GB. Adenosine receptor activation in brain reduces stress-induced ulcer
formation. Eur J Pharmacol 1985;115:185–190. [PubMed: 2998819]
Cunha et al. Page 20
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. [PubMed:
14636179]
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. [PubMed: 2812276]
168. Herman JP, Cullinan WE. Neurocircuitry of stress: central control of the hypothalamo-pituitary-
adrenocortical axis. Trends Neurosci 1997;20:78–84. [PubMed: 9023876]
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. [PubMed: 10525758]
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. [PubMed: 15718055]
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. [PubMed: 11160442]
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. [PubMed: 12384217]
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. [PubMed: 9405958]
174. McEwen BS, Magarinos AM, Reagan LP. Structural plasticity and tianeptine: cellular and molecular
targets. Eur Psychiatry 2002;17(Suppl 3):318–330. [PubMed: 15177088]
175. Gillman PK. Tricyclic antidepressant pharmacology and therapeutic drug interactions updated. Br
J Pharmacol 2007;151:737–748. [PubMed: 17471183]
176. Jeanneteau F, Chao MV. Promoting neurotrophic effects by GPCR ligands. Novartis Found Symp
2006;276:181–189. [PubMed: 16805430]
177. Martinowich K, Manji H, Lu B. New insights into BDNF function in depression and anxiety. Nature
Neurosci 2007;10:1089–1093. [PubMed: 17726474]
178. Martinowich K, Lu B. Interaction between BDNF and serotonin: role in mood disorders.
Neuropsychopharmacology 2008;33:73–83. [PubMed: 17882234]
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. [PubMed: 15044529]
180. Lee FS, Chao MV. Activation of Trk neurotrophin receptors in the absence of neurotrophins. Proc
Natl Acad Sci USA 2001;98:3555–3560. [PubMed: 11248116]
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. [PubMed: 18005343]
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. [PubMed: 16957081]
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. [PubMed: 14729225]
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. [PubMed: 16202183]
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. [PubMed: 17868670]
Cunha et al. Page 21
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
186. Lara DR, Souza DO. Schizophrenia: a purinergic hypothesis. Med Hypotheses 2000;54:157–166.
[PubMed: 10790742]
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. [PubMed: 16580767]
188. De Freitas B, Schwartz G. Effects of caffeine in chronic psychiatric patients. Am J Psychiatry
1979;10:1337–1338. [PubMed: 484737]
189. Mikkelsen EJ. Caffeine and schizophrenia. J Clin Psychiatry 1978;9:732–736. [PubMed: 690092]
190. Nickell PV, Uhde TW. Dose–response effects of intravenous caffeine in normal volunteers. Anxiety
1994;1:161–168. [PubMed: 9160568]
191. Mayo KM, Falkowski W, Jones CA. Caffeine: use and effects in long-stay psychiatric patients. Br
J Psychiatry 1993;162:543–545. [PubMed: 8481748]
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. [PubMed: 10849191]
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. [PubMed: 15694232]
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.
[PubMed: 11459338]
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. [PubMed: 2375945]
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. [PubMed:
16469942]
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. [PubMed: 12377399]
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. [PubMed: 9596047]
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. [PubMed: 7612156]
200. Hyde TM, Weinberger DR. Seizures and schizophrenia. Schizophr Bull 1997;23:611–622.
[PubMed: 9365998]
201. Dunwiddie TV. Adenosine and suppression of seizures. Adv Neurol 1999;79:1001–1010. [PubMed:
10514882]
202. Rivkees SA, Zhao Z, Porter G, Turner C. Influences of adenosine on the fetus and newborn. Mol
Genet Metab 2001;74:160–171. [PubMed: 11592813]
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. [PubMed:
17044013]
204. Othman T, Yan H, Rivkees SA. Oligodendrocytes express functional A1 adenosine receptors that
stimulate cellular migration. Glia 2003;44:166–172. [PubMed: 14515332]
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. [PubMed:
12467589]
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. [PubMed:
14960625]
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. [PubMed: 10926292]
Cunha et al. Page 22
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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.
209. Rivkees SA. The ontogeny of cardiac and neural A1 adenosine receptor expression in rats. Dev
Brain Res 1995;89:202–213. [PubMed: 8612324]
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. [PubMed: 12827647]
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.
[PubMed: 12900046]
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. [PubMed: 18005073]
213. Ferré S. Adenosine–dopamine interactions in the ventral striatum. Implications for the treatment of
schizophrenia. Psychopharmacology 1997;2:107–20.
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. [PubMed: 10884434]
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. [PubMed: 16786561]
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.
[PubMed: 9739141]
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. [PubMed: 17254024]
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.
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. [PubMed: 9885784]
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. [PubMed: 11490180]
221. Olney JW, Farber NB. Glutamate receptor dysfunction and schizophrenia. Arch Gen Psychiatry
1995;12:998–1007. [PubMed: 7492260]
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.
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. [PubMed: 10225370]
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.
[PubMed: 1354544]
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. [PubMed: 7679722]
226. Manzoni OJ, Manabe T, Nicoll RA. Release of adenosine by activation of NMDA receptors in the
hippocampus. Science 1994;265:2098–2101. [PubMed: 7916485]
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: 9580599]
Cunha et al. Page 23
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. [PubMed: 10594340]
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. [PubMed: 11986374]
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.
[PubMed: 16809361]
231. Moghaddam B, Adams BW. Reversal of phencyclidine effects by a group II metabotropic glutamate
receptor agonist in rats. Science 1998;281:1349–1352. [PubMed: 9721099]
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. [PubMed: 12700703]
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.
234. Rimondini R, Ferré S, Ogren SO, Fuxe K. Adenosine A2A agonists: a potential new type of atypical
antipsychotic. Neuropsychopharmacology 1997;2:82–91. [PubMed: 9252983]
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. [PubMed: 7662885]
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. [PubMed: 12231386]
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.
[PubMed: 8618915]
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. [PubMed: 16271052]
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.
[PubMed: 15081796]
240. Hughes JR, McHugh P, Holtzman S. Caffeine and schizophrenia. Psychiatr Serv 1998;11:1415–
1417. [PubMed: 9826240]
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. [PubMed: 9767181]
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. [PubMed: 12634474]
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. [PubMed:
9029412]
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. [PubMed: 15719154]
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.
Cunha et al. Page 24
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. p. 1090-1182.
247. Biederman J, Faraone SV. Attention-deficit hyperactivity disorder. Lancet 2005;366:237–248.
[PubMed: 16023516]
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. [PubMed:
15950004]
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. [PubMed: 17318414]
250. Wilens TE. Attention deficit hyperactivity disorder and substance use disorders. Am J Psychiatry
2006;163:2059–2063. [PubMed: 17151154]
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.
[PubMed: 16546312]
252. Arnsten AF. Fundamentals of attention-deficit/hyperactivity disorder: circuits and pathways. J Clin
Psychiatry 2006;67(Suppl 8):7–12. [PubMed: 16961424]
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. [PubMed: 16894395]
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. [PubMed: 12808433]
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. [PubMed: 11431226]
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. [PubMed: 16774975]
257. Oak JN, Oldenhof J, Van Tol HH. The dopamine D4 receptor: one decade of research. Eur J
Pharmacol 2000;405:303–327. [PubMed: 11033337]
258. Gross MD. Caffeine in the treatment of children with minimal brain dysfunction or hyperkinetic
syndrome. Psychosomatics 1975;16:26–27. [PubMed: 1101283]
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. [PubMed: 7028238]
260. Arnsten AF. Stimulants: Therapeutic actions in ADHD. Neuropsychopharmacology 2006;31:2376–
2383. [PubMed: 16855530]
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.
[PubMed: 1096645]
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: 3072303]
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. [PubMed:
1896253]
Cunha et al. Page 25
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
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. [PubMed:
7696282]
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. [PubMed: 10197295]
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. [PubMed: 8603222]
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. [PubMed:
12204387]
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. [PubMed: 17241669]
269. Jentsch JD. Impaired visuospatial divided attention in the spontaneously hypertensive rat. Behav
Brain Res 2005;157:323–330. [PubMed: 15639183]
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.
[PubMed: 17395336]
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. [PubMed: 11844582]
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. [PubMed: 15877934]
Cunha et al. Page 26
Curr Pharm Des. Author manuscript; available in PMC 2009 January 1.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
... Low basal extracellular levels of adenosine [56] and likely a slight increase in adenosine levels preferentially activate neuronal A1Rs, resulting in inhibitory effects in the central nervous system, such as a decrease in glutamate release [57]. Nevertheless, a higher increase in adenosine levels can activate A2A-type adenosine receptors (A2ARs) in the A1R-A2AR heteromer, blocking the functions of A1Rs and generating enhanced glutamate release [57,58]. Thus, it is possible that administration of higher doses of ketone supplements (e.g., 5 g/kg KEMCT/day) could generate higher levels of not only blood R-βHB ( Figure 1C) but also adenosine compared with the effects evoked by lower doses of ketone supplements, in which higher adenosine levels may be enough for the preferential activation of neuronal A2ARs [59]. ...
Article
Full-text available
While one-third of the population can be affected by anxiety disorders during their lifetime, our knowledge of the pathophysiology of these disorders is far from complete. Previously, it has been demonstrated in male animals that exogenous ketone supplement-evoked ketosis can decrease anxiety levels in preclinical rodent models, such as Wistar Albino Glaxo/Rijswijk (WAG/Rij) rats. Thus, in this study, we investigated whether intragastric gavage of the exogenous ketone supplement KEMCT (mix of 1,3-butanediol-acetoacetate diester/ketone ester/KE and medium-chain triglyceride/MCT oil in 1:1 ratio) for 7 days can alter the anxiety levels of female WAG/Rij rats using the light-dark box (LDB) test. We demonstrated that a lower dose of KEMCT (3 g/kg/day) increased blood R-βHB (R-β-hydroxybutyrate) levels and significantly decreased anxiety levels (e.g., increased the time spent in the light compartment) in female WAG/Rij rats on the seventh day of administration. Although the higher KEMCT dose (5 g/kg/day) increased blood R-βHB levels more effectively, compared with the lower KEMCT dose, anxiety levels did not improve significantly. We conclude that ketone supplementation might be an effective strategy to induce anxiolytic effects not only in male but also in female WAG/Rij rats. However, these results suggest that the optimal level may be moderately, not highly, elevated blood R-βHB levels when the goal is to alleviate symptoms of anxiety. More studies are needed to understand the exact mechanism of action of ketone supplementation on anxiety levels and to investigate their use in other animal models and humans for the treatment of anxiety disorders and other mental health conditions.
... It also plays a critical role in other biological processes, including the regulation of mood and behavior, by acting on various receptors such as the A 1 , A 2A , A 2B , and A 3A subtypes (Pasquini et al., 2022). It is reported that adenosine and its analogs produced depressant-like behavioral effects in animal models of depression (Woodson et al., 1998;Hunter et al., 2003;Cunha et al., 2008). Thus, elevated adenosine levels extended the immobilization time in rats subjected to inescapable shocks as well as in the FST. ...
Article
Full-text available
Introduction: Post-traumatic stress disorder (PTSD) is a chronic mental illness triggered by traumatic experiences such as wars, natural disasters, or catastrophes, and it is characterized by anxiety, depression and cognitive impairment. Diosgenin is a steroidal sapogenin with known neuroprotective and antioxidant properties. This study aimed to assess the pharmacological potential of diosgenin in a single prolonged stress (SPS) model of PTSD, plus other behavioral models along with any consequent alterations in brain neurochemistry in male mice. Methodology: SPS was induced by restraining animals for 2 h, followed by 20 min of forced swim, recuperation for 15 min, and finally, exposure to ether to induce anesthesia. The SPS-exposed animals were treated with diosgenin (20, 40, and 60 mg/kg) and compared with the positive controls, fluoxetine or donepezil, then they were observed for any changes in anxiety/depression-like behaviors, and cognitive impairment. After behavioral screening, postmortem serotonin, noradrenaline, dopamine, vitamin C, adenosine and its metabolites inosine and hypoxanthine were quantified in the frontal cortex, hippocampus, and striatum by high-performance liquid chromatography. Additionally, animal serum was screened for changes in corticosterone levels. Results: The results showed that diosgenin reversed anxiety- and depression-like behaviors, and ameliorated cognitive impairment in a dose-dependent manner. Additionally, diosgenin restored monoamine and vitamin C levels dose-dependently and modulated adenosine and its metabolites in the brain regions. Diosgenin also reinstated otherwise increased serum corticosterone levels in SPS mice. Conclusion: The findings suggest that diosgenin may be a potential candidate for improving symptoms of PTSD.
... The association of the adenosergic system with mood disorders has primarily been established, employing adenosine and its analogs to cause depression-like behavioral effects in extensively used animal models of depression [59][60][61]. Thus, an upsurge in adenosine levels prolonged the immobilization time in rats presented with inescapable shocks and in the FST [62,63]. ...
... Intriguingly, the latter finding was not supported in the Asian population [85], which raises the possibility that the effect is ethnically unique. As a result of the finding that D2R antagonists have antipsychotic effects and the antagonistic relationship between A 2A and D 2 receptors, A 2A Rs have also drawn interest as a possible target for the treatment of schizophrenia [86][87][88][89]. ...
Article
Full-text available
All biological tissues and bodily fluids include the autacoid adenosine. The P1 class of purinergic receptors includes adenosine receptors. Four distinct G-protein-coupled receptors on the cellular membrane mediate the effects of adenosine, whose cytoplasmic content is regulated by producing/degrading enzymes and nucleoside transporters. A2A receptor has received a great deal of attention in recent years because it has a wide range of potential therapeutic uses. A2B and, more significantly, A2A receptors regulate numerous physiological mechanisms in the central nervous system (CNS). The inferior targetability of A2B receptors towards adenosine points that they might portray a promising medicinal target since they are triggered only under pharmacological circumstances (when adenosine levels rise up to micromolar concentrations). The accessibility of specific ligands for A2B receptors would permit the exploration of such a theory. A2A receptors mediate both potentially neurotoxic and neuroprotective actions. Hence, it is debatable to what extent they play a role in neurodegenerative illnesses. However, A2A receptor blockers have demonstrated clear antiparkinsonian consequences, and a significant attraction exists in the role of A2A receptors in other neurodegenerative disorders. Amyloid peptide extracellular accumulation and tau hyperphosphorylation are the pathogenic components of AD that lead to neuronal cell death, cognitive impairment, and memory loss. Interestingly, in vitro and in vivo research has shown that A2A adenosine receptor antagonists may block each of these clinical symptoms, offering a crucial new approach to combat a condition for which, regrettably, only symptomatic medications are currently available. At least two requirements must be met to determine whether such receptors are a target for diseases of the CNS: a complete understanding of the mechanisms governing A2A-dependent processes and the availability of ligands that can distinguish between the various receptor populations. This review concisely summarises the biological effects mediated by A2A adenosine receptors in neurodegenerative disorders and discusses the chemical characteristics of A2A adenosine receptor antagonists undergoing clinical trials. Graphical Abstract Selective A2A receptor blocker against neurodegenerative disorders
... Finally, we chose to target the adenosine receptor A2a (ADORA2A) as a second exemplar membrane target. This GPCR has become an important potential target for immunotherapy (55), as well as psychiatric and degenerative disorders (56), but critically, has never been identified through live cell chemoproteomics (57,58). Using a reported ligand for ADORA2A ...
Article
Full-text available
Over half of new therapeutic approaches fail in clinical trials due to a lack of target validation. As such, the development of new methods to improve and accelerate the identification of cellular targets, broadly known as target ID, remains a fundamental goal in drug discovery. While advances in sequencing and mass spectrometry technologies have revolutionized drug target ID in recent decades, the corresponding chemical-based approaches have not changed in over 50 y. Consigned to outdated stoichiometric activation modes, modern target ID campaigns are regularly confounded by poor signal-to-noise resulting from limited receptor occupancy and low crosslinking yields, especially when targeting low abundance membrane proteins or multiple protein target engagement. Here, we describe a broadly general platform for photocatalytic small molecule target ID, which is founded upon the catalytic amplification of target-tag crosslinking through the continuous generation of high-energy carbene intermediates via visible light-mediated Dexter energy transfer. By decoupling the reactive warhead tag from the small molecule ligand, catalytic signal amplification results in unprecedented levels of target enrichment, enabling the quantitative target and off target ID of several drugs including (+)-JQ1, paclitaxel (Taxol), dasatinib (Sprycel), as well as two G-protein-coupled receptors-ADORA2A and GPR40.
Article
Full-text available
Background Depression is a prevalent psychiatric disorder with high long‐term morbidities, recurrences, and mortalities. Despite extensive research efforts spanning decades, the cellular and molecular mechanisms of depression remain largely unknown. What's more, about one third of patients do not have effective anti‐depressant therapies, so there is an urgent need to uncover more mechanisms to guide the development of novel therapeutic strategies. Adenosine triphosphate (ATP) plays an important role in maintaining ion gradients essential for neuronal activities, as well as in the transport and release of neurotransmitters. Additionally, ATP could also participate in signaling pathways following the activation of postsynaptic receptors. By searching the website PubMed for articles about “ATP and depression” especially focusing on the role of extracellular ATP (eATP) in depression in the last 5 years, we found that numerous studies have implied that the insufficient ATP release from astrocytes could lead to depression and exogenous supply of eATP or endogenously stimulating the release of ATP from astrocytes could alleviate depression, highlighting the potential therapeutic role of eATP in alleviating depression. Aim Currently, there are few reviews discussing the relationship between eATP and depression. Therefore, the aim of our review is to conclude the role of eATP in depression, especially focusing on the evidence and mechanisms of eATP in alleviating depression. Conclusion We will provide insights into the prospects of leveraging eATP as a novel avenue for the treatment of depression.
Article
KW‐6356 is a selective antagonist and inverse agonist of the adenosine A 2A receptor. The primary aim of the present analysis was to characterize the pharmacokinetics (PK) of KW‐6356 and its active metabolite M6 in healthy subjects and patients with Parkinson's disease (PD). We pooled concentration‐time data from healthy subjects and patients with PD who were administered KW‐6356. Using these data, we developed a population PK model by sequentially fitting the KW‐6356 parameters followed by the M6 parameters. A first‐order absorption with a 1‐compartment model for KW‐6356 and a 1‐compartment model for M6 best described the profiles. The covariates included in the final models were food status (fed/fasted/unknown) on first‐order absorption rate constant, baseline serum albumin level on apparent clearance of KW‐6356, and baseline body weight on apparent volume of distribution of KW‐6356 and apparent clearance of M6. No covariate had a clinically meaningful impact on KW‐6356 or M6 exposure.
Article
KW‐6356 is an adenosine A2A receptor–selective antagonist and inverse agonist. We conducted 2 studies: study 6356‐001 (no NCT number), a randomized, double‐blind, placebo‐controlled phase 1 trial of single ascending (1, 3, 10 mg) and multiple (6 mg once daily) oral doses of KW‐6356 in healthy Japanese subjects; and study 6356‐004 (NCT03830528), including a randomized, double‐blind, placebo‐controlled phase 1 trial of single ascending (21, 42, 60 mg) and multiple (24 mg once daily) oral doses of KW‐6356, and a phase 1 open‐label trial of multiple oral doses (6 mg once daily) of KW‐6356 in healthy Japanese and White subjects, to evaluate the safety, tolerability, and pharmacokinetics of KW‐6356. KW‐6356 was well tolerated after administration of single doses of up to 60 mg and multiple doses of up to 24 mg once daily for 14 days. The pharmacokinetics of KW‐6356 were linear after a single dose of up to 60 mg KW‐6356. The mean terminal elimination half‐life of KW‐6356 ranged from 18.4 to 43.1 hours following administration of single doses of 1–60 mg. There was no clear difference in the safety or pharmacokinetics of KW‐6356 between healthy Japanese and White subjects.
Article
Anabolic-androgenic steroids (AAS) and caffeine can induce several behavioral alterations in humans and rodents. Administration of nandrolone decanoate is known to affect defensive responses to aversive stimuli, generally decreasing inhibitory control and increasing aggressivity but whether caffeine intake influences behavioral changes induced by AAS is unknown. The present study aimed to investigate behavioral effects of caffeine (a non-selective antagonist of adenosine receptors) alone or combined with nandrolone decanoate (one of the most commonly AAS abused) in female and male Lister Hooded rats. Our results indicated that chronic administration of nandrolone decanoate (10 mg/kg, i.m., once a week for 8 weeks) decreased risk assessment/anxiety-like behaviors (in the elevated plus maze test), regardless of sex. These effects were prevented by combined caffeine intake (0.1 g/L, p.o., ad libitum). Overall, the present study heralds a key role for caffeine intake in the modulation of nandrolone decanoate-induced behavioral changes in rats, suggesting adenosine receptors as candidate targets to manage impact of AAS on brain function and behavior.
Article
Full-text available
Extracellular adenosine critically modulates ischemic brain injury, at least in part through activation of the A 1 adenosine receptor. However, the role played by the A 2A receptor has been obscured by intrinsic limitations of A 2A adenosinergic agents. To overcome these pharmacological limitations, we explored the consequences of deleting the A 2A adenosine receptor on brain damage after transient focal ischemia. Cerebral morphology, as well as vascular and physiological measures (before, during, and after ischemia) did not differ between A 2A receptor knock-out and wild-type littermates. The volume of cerebral infarction, as well as the associated neurological deficit induced by transient filament occlusion of the middle cerebral artery, were significantly attenuated in A 2A receptor knock-out mice. This neuroprotective phenotype of A 2A receptor-deficient mice was observed in different genetic backgrounds, confirming A 2A receptor disruption as its cause. Together with complimentary pharmacological studies, these data suggest that A 2A receptors play a prominent role in the development of ischemic injury within brain and demonstrate the potential for anatomical and functional neuroprotection against stroke by A 2A receptor antagonists.
Article
Full-text available
New data on the pharmacology of tricyclic antidepressants (TCAs), their affinities for human cloned CNS receptors and their cytochrome P450 enzyme inhibition profiles, allow improved deductions concerning their effects and interactions and indicate which of the TCAs are the most useful. The relative toxicity of TCAs continues to be more precisely defined, as do TCA interactions with selective serotonin reuptake inhibitors (SSRIs). TCA interactions with monoamine oxidase inhibitors (MAOIs) have been, historically, an uncertain and difficult question, but are now well understood, although this is not reflected in the literature. The data indicate that nortriptyline and desipramine have the most pharmacologically desirable characteristics as noradrenaline reuptake inhibitors (NRIs), and as drugs with few interactions that are also safe when coadministered with either MAOIs or SSRIs. Clomipramine is the only available antidepressant drug that has good evidence of clinically relevant serotonin and noradrenaline reuptake inhibition (SNRI). These data assist drug selection for monotherapy and combination therapy and predict reliably how and why pharmacodynamic and pharmacokinetic interactions occur. In comparison, two newer drugs proposed to have SNRI properties, duloxetine and venlafaxine, may have insufficient NRI potency to be effective SNRIs. Combinations such as sertraline and nortriptyline may therefore offer advantages over drugs like venlafaxine that have fixed ratios of SRI/NRI effects that are not ideal. However, no TCA/SSRI combination is sufficiently safe to be universally applicable without expert knowledge. Standard texts (e.g. the British National Formulary) and treatment guidelines would benefit by taking account of these new data and understandings. British Journal of Pharmacology (2007) 151, 737–748; doi:10.1038/sj.bjp.0707253
Article
Full-text available
The treatment guideline draws on several international guidelines: ( i ) Practice Guidelines of the American Psychiatric Association (APA) for the Treatment of Patients with Major Depressive Disorder, Second Edition; [1] ( ii ) Clinical Guidelines for the Treatment of Depressive Disorders by the Canadian Psychiatric Association and the Canadian Network for Mood and Anxiety Treatments (CANMAT); [2] ( iii ) National Institute for Clinical Excellence (NICE) guidelines; [3] ( iv ) Royal Australian and New Zealand College of Psychiatrists Clinical Practice Guidelines Team for Depression (RANZCAP); [4] ( v ) Texas Medication Algorithm Project (TMAP) Guidelines; [5] ( vi ) World Federation of Societies of Biological Psychiatry (WFSBP) Treatment Guideline for Unipolar Depressive Disorder; [6] and ( vii ) British Association for Psychopharmacology Guidelines. [7
Article
Attention deficit hyperactivity disorder (ADHD) is a common neurodevelopmental disorder, starting in early childhood and persisting into adulthood in the majority of cases. Family and twin studies have demonstrated the importance of genetic factors and candidate gene association studies have identified several loci that exert small but significant effects on ADHD. To provide further clarification of reported associations and identify novel associated genes, we examined 1038 single-nucleotide polymorphisms (SNPs) spanning 51 candidate genes involved in the regulation of neurotransmitter pathways, particularly dopamine, norepinephrine and serotonin pathways, in addition to circadian rhythm genes. Analysis used within family tests of association in a sample of 776 DSM-IV ADHD combined type cases ascertained for the International Multi-centre ADHD Gene project. We found nominal significance with one or more SNPs in 18 genes, including the two most replicated findings in the literature: DRD4 and DAT1. Gene-wide tests, adjusted for the number of SNPs analysed in each gene, identified associations with TPH2, ARRB2, SYP, DAT1, ADRB2, HES1, MAOA and PNMT. Further studies will be needed to confirm or refute the observed associations and their generalisability to other samples.
Article
Objective: To investigate the possible association of dietary caffeine consumption and medicinal caffeine use with chronic daily headache (CDH). Methods: Population-based cases and controls were recruited from the Baltimore, MD, Philadelphia, PA, and Atlanta, GA, metropolitan areas. Controls ( n = 507) reported 2 to 104 headache days/year, and cases ( n = 206) reported greater than or equal to 180 headache days/year. Current and past dietary caffeine consumption and medication use for headache were based on detailed self-report. High caffeine exposure was defined as being in the upper quartile of dietary consumption or using a caffeine-containing over-the-counter analgesic as the preferred headache treatment. Results: In comparison with episodic headache controls, CDH cases were more likely overall to have been high caffeine consumers before onset of CDH ( odds ratio [ OR] = 1.50, p = 0.05). No association was found for current caffeine consumption (i.e., post CDH) ( OR = 1.36, p = 0.12). In secondary analyses, associations were confined to younger ( age < 40) women ( OR = 2.0, p = 0.02) and those with chronic episodic ( as opposed to chronic continuous) headaches ( OR = 1.69, p = 0.01), without physician consultation ( OR = 1.67, p = 0.04) and of recent ( < 2 years) onset ( OR = 1.67, p = 0.03). Conclusion: Dietary and medicinal caffeine consumption appears to be a modest risk factor for chronic daily headache onset, regardless of headache type.
Article
Objective: There is growing interest in investigating the adenosine-dopamine interaction in the ventral striatum. Adenosine plays a role opposite to dopamine in the striatum and adenosine antagonists, like caffeine, produce similar effects to increased dopaminergic neurotransmission in the striatum. In particular, a strong antagonistic interaction between adenosine A2A and dopamine D2 receptors takes place in the striopallidal GABAergic neurones. Therefore, adenosine agonists or uptake inhibitors provide a potential new treatment for schizophrenia. We undertook a pilot trial to investigate whether the combination of haloperidol with dipyridamole, an uptake inhibitor of adenosine, was more effective than haloperidol alone. Methods: Thirty patients who met the DSM IV criteria for schizophrenia completed the study. Patients were allocated in a random fashion, 16 to haloperidol 20 mg/day plus dipyridamole 75 mg/day and 14 to haloperidol 20 mg/day plus placebo. Results: Although both protocols significantly decreased the score of the positive, negative and general psychopathological symptoms over the trial period, the combination of haloperidol and dipyridamole was significantly better than haloperidol alone in decreasing positive and general psychopathology symptoms as well as PANSS total scores. Conclusion: Dipyridamole may be of therapeutic benefit in treating schizophrenia in combination with neuroleptics. However, a larger study to confirm our results is warranted.
Article
Adenosine is a neuromodulator in the nervous system and it has recently been observed that pharmacological blockade or gene disruption of adenosine A(2A) receptors confers neuroprotection under different neurotoxic situations in the brain. We now observed that coapplication of either caffeine (1-25 mum) or the selective A(2A) receptor antagonist, 4-(2-[7-amino-2(2-furyl)(1,2,4)triazolo (2,3-a)(1,3,5)triazin-5-ylamino]ethyl)phenol (ZM 241385, 50 nm), but not the A receptor antagonist, 8-cyclopentyltheophylline (200 nm), prevented the neuronal cell death caused by exposure of rat cultured cerebellar granule neurons to fragment 25 - 35 of beta-amyloid protein (25 mum for 48 h), that by itself caused a near three-fold increase of propidium iodide-labeled cells. This constitutes the first in vitro evidence to suggest that adenosine A(2A) receptors may be the molecular target responsible for the observed beneficial effects of caffeine consumption in the development of Alzheimer's disease.
Article
Caffeine is the most widely consumed psychostimulant drug in the world. With intermittent exposures, caffeine may act as a mild analgesic for headache or as an adjuvant for the actions of other analgesics. Chronic repetitive exposures to caffeine increase the risks for development of analgesic-overuse headache, chronic daily headache, and physical dependency. Cessation of caffeine use after chronic exposures leads to a withdrawal syndrome with headache as a dominant symptom. At dosages achieved by common dietary intake, caffeine acts as a potent antagonist of central and peripheral nervous system adenosine receptors. The complex effects of caffeine on headache disorders suggest important roles for adenosine in these disorders and in the induction of caffeine dependency.