ArticlePDF Available

Dopamine and glutamate in schizophrenia: biology, symptoms and treatment

Authors:

Abstract and Figures

Glutamate and dopamine systems play distinct roles in terms of neuronal signalling, yet both have been proposed to contribute significantly to the pathophysiology of schizophrenia. In this paper we assess research that has implicated both systems in the aetiology of this disorder. We examine evidence from post‐mortem, preclinical, pharmacological and in vivo neuroimaging studies. Pharmacological and preclinical studies implicate both systems, and in vivo imaging of the dopamine system has consistently identified elevated striatal dopamine synthesis and release capacity in schizophrenia. Imaging of the glutamate system and other aspects of research on the dopamine system have produced less consistent findings, potentially due to methodological limitations and the heterogeneity of the disorder. Converging evidence indicates that genetic and environmental risk factors for schizophrenia underlie disruption of glutamatergic and dopaminergic function. However, while genetic influences may directly underlie glutamatergic dysfunction, few genetic risk variants directly implicate the dopamine system, indicating that aberrant dopamine signalling is likely to be predominantly due to other factors. We discuss the neural circuits through which the two systems interact, and how their disruption may cause psychotic symptoms. We also discuss mechanisms through which existing treatments operate, and how recent research has highlighted opportunities for the development of novel pharmacological therapies. Finally, we consider outstanding questions for the field, including what remains unknown regarding the nature of glutamate and dopamine function in schizophrenia, and what needs to be achieved to make progress in developing new treatments.
Content may be subject to copyright.
SPECIAL ARTICLE
World Psychiatry 19:1 - February 2020 15
Dopamine and glutamate in schizophrenia: biology, symptoms
and treatment
Robert A. McCutcheon1-3, John H. Krystal4-6, Oliver D. Howes1-3
1Institute of Psychiatry, Psychology and Neuroscience, King’s College London, London, UK; 2MRC London Institute of Medical Sciences, Imperial College London, Hammer-
smith Hospital, London, UK; 3South London and Maudsley Foundation NHS Trust, Maudsley Hospital, London, UK; 4Department of Radiology and Biomedical Imaging, Yale
University School of Medicine, New Haven, CT, USA; 5Department of Psychiatry, Yale University School of Medicine, New Haven, CT, USA; 6VA National Center for PTSD,
VA Connecticut Healthcare System, West Haven, CT, USA
Glutamate and dopamine systems play distinct roles in terms of neuronal signalling, yet both have been proposed to contribute signicantly to
the pathophysiology of schizophrenia. In this paper we assess research that has implicated both systems in the aetiology of this disorder. Weex-
amine evidence from post-mortem, preclinical, pharmacological and in vivo neuroimaging studies. Pharmacological and preclinical studies
implicate both systems, and in vivo imaging of the dopamine system has consistently identied elevated striatal dopamine synthesis and release
capacity in schizophrenia. Imaging of the glutamate system and other aspects of research on the dopamine system have produced less consistent
ndings, potentially due to methodological limitations and the heterogeneity of the disorder. Converging evidence indicates that genetic and
environmental risk factors for schizophrenia underlie disruption of glutamatergic and dopaminergic function. However, while genetic inuences
may directly underlie glutamatergic dysfunction, few genetic risk variants directly implicate the dopamine system, indicating that aberrant do-
pamine signalling is likely to be predominantly due to other factors. We discuss the neural circuits through which the two systems interact, and
how their disruption may cause psychotic symptoms. We also discuss mechanisms through which existing treatments operate, and how recent
research has highlighted opportunities for the development of novel pharmacological therapies. Finally, we consider outstanding questions for
the eld, including what remains unknown regarding the nature of glutamate and dopamine function in schizophrenia, and what needs to be
achieved to make progress in developing new treatments.
Key words: Psychosis, schizophrenia, dopamine, glutamate, antipsychotics, striatum, NMDA receptors, D2 receptors, D1 receptors, dorsolateral
prefrontal cortex, GABA interneurons, amphetamine, ketamine, cognitive symptoms
(World Psychiatry 2020;19:15–33)
Schizophrenia is a severe mental disorder characterized by
positive symptoms such as delusions and hallucinations, nega-
tive symptoms including amotivation and social withdrawal,
and cognitive symptoms such as decits in working memory and
cognitive exibility1. e disorder accounts for signicant health
care costs, and is associated with a reduced life expectancy of
about 15 years on average2.
Antipsychotics were serendipitously discovered over fty years
ago, but it took another decade or so until dopamine antagonism
was demonstrated as central to their clinical eectiveness3. Fur-
ther evidence implicating the dopamine system in the patho-
physiology of schizophrenia has subsequently accumulated, and
it remains the case that all licensed rst-line treatments for schiz-
ophrenia operate primarily via antagonism of the dopamine D2
receptor4.
However, despite the central role that dopamine plays in our
understanding of schizophrenia, it has also become increasingly
clear that dysfunction of this system may not be sucient to ex-
plain several phenomena. In particular, dopamine blockade is
not an eective treatment for negative and cognitive symptoms
and, in a signicant proportion of patients, it does not improve
positive symptoms either. As a result, attention has turned to ad-
ditional neurochemical targets. Glutamate is the major excita-
tory neurotransmitter of the central nervous system. e nding
that antagonists of a specic glutamate receptor, the N-methyl-
D-aspartate (NMDA) receptor, induce psychotic symptoms has
led to a wealth of research implicating the glutamate system in
the pathophysiology of schizophrenia.
In this paper we review the evidence regarding dopaminer-
gic and glutamatergic functioning in schizophrenia. We survey
indirect ndings from preclinical, genetic and pharmacologi-
cal studies, evidence from post-mortem research, and results of
neuroimaging studies that characterize functioning in living pa-
tients. We discuss how dysregulation of these systems may lead
to the symptoms of the disorder, and the therapeutic possibilities
associated with their pharmacological modulation. We then ex-
plore what may underlie this dysregulation, and the interaction
between the two systems, before concluding by considering out-
standing questions for the eld.
DOPAMINE
Dopamine was initially thought to be a biologically inactive
intermediary compound on the synthetic pathway between tyro-
sine and noradrenaline. Work by A. Carlsson and others, however,
demonstrated that dopamine depletion inhibited movement,
and that this eect could be reversed following the administra-
tion of the dopamine precursor L-DOPA. is established that
the molecule was of major biological importance in its own right5,
and discrete dopaminergic projections were subsequently identi-
ed.
at dopaminergic dysfunction might play a role in the de-
velopment of psychotic symptoms is one of the longest stand-
ing hypotheses regarding the pathophysiology of schizophrenia.
Below, we discuss the evidence for dopamine dysfunction in
16 World Psychiatry 19:1 - Februar y 2020
schizophrenia, before considering how this may lead to psychot-
ic symptoms, and the mechanisms through which dopamine
modulating treatments exert their eects.
Indirect evidence for dopamine dysfunction in
schizophrenia
Animal models
Rodent models of schizophrenia are useful for investigating
molecular mechanisms that may be of pathophysiological rel-
evance, and for testing novel therapeutic interventions.
One well characterized model of dopaminergic hyperactiv-
ity involves administering repeated doses of amphetamine. is
has been shown to induce events that are also observed in indi-
viduals with schizophrenia, such as reduced prepulse inhibition,
stereotyped behaviours, and impaired cognitive exibility and
attention6. Given that amphetamine results in dopamine release,
and that the above eects can be ameliorated with the adminis-
tration of dopamine antagonists, this provides indirect evidence
for a role of dopamine in behaviour thought to be a proxy for psy-
chotic symptoms.
Another example is that of mice genetically modied to over-
express dopamine D2 receptors in the striatum, which also dis-
play a wide range of schizophrenia-like behaviours7. Similarly,
transgenic insertion of tyrosine hydroxylase and guanosine tri-
phosphate (GTP) cyclohydrase 1 into the substantia nigra in
early adolescence increases dopamine synthesis capacity, and
has been associated with a schizophrenia-like behavioural phe-
notype8.
Other examples do not target the dopamine system directly,
but are still associated with dopaminergic abnormalities. e
methylazoxymethanol acetate (MAM) model involves inducing
neurodevelopmental disruption of the hippocampus via the ad-
ministration of MAM to pregnant rats, and is accompanied by in-
creased ring rates of mesostriatal dopamine neurons9. A model
of environmental risk factors in which rats were socially isolated
post weaning has also been associated with increased striatal pre -
synaptic dopamine function10.
In summary, multiple methods have been used to induce in-
creased striatal dopamine signalling in animal models, and these
consistently produce behaviours analogous to those observed in
individuals with schizophrenia.
Cerebrospinal fluid and post-mortem studies
Studies examining levels of dopamine and its metabolites
– 3,4-dihydroxyphenylacetic acid (DOPAC) and homovanillic
acid (HVA) – in schizophrenia, both peripherally and in cere-
brospinal uid (CSF), have given inconsistent results11-13. is
may be due to the fact that these levels are a state dependent
marker, and to the eects of antipsychotic treatment. Studies
have found that levels of dopamine, HVA and DOPAC in CSF are
only increased in those receiving antipsychotic treatment13,14,
and that reductions occur following the withdrawal of antipsy-
chotics15,16.
Some17-19, but not all20, studies of HVA have found higher lev-
els in both CSF and plasma of acutely relapsed patients com-
pared to stable patients. ere have also been suggestions that
baseline plasma HVA levels may predict subsequent response to
antipsychotics, which shows some parallels with imaging nd-
ings considered below21.
is approach to studying the dopamine system, however, has
declined in popularity over recent years. A major weakness is
that the measurement occurs distal from the dopamine neurons
of interest. Since both hypo- and hyperdopaminergic function
may exist within an individual simultaneously, a technique that
allows for anatomical specicity is required to understand the
nature and localization of changes.
Early post-mortem investigations suggested that striatal D2
receptor levels might be raised in individuals with psychosis22,
and a meta-analysis of seven post-mortem studies suggested
that receptor levels were increased with a large eect size23. How-
ever, no studies of antipsychotic naïve individuals exist, and
themajority are of individuals chronically treated with antipsy-
chotic medications, which have been found to lead to D2 recep-
tor upregulation24,25.
Post-mortem studies have also examined the substantia nigra.
In these studies evidence regarding dopamine function is incon-
sistent, with some studies suggesting an increase in tyrosine hy-
droxylase levels in patients26, but others nding no dierence27,28.
Other studies have found abnormal nuclear morphology of sub-
stantia nigra neurons29, reduced dopamine transporter (DAT)
and vesicular monoamine transporter (VMAT) gene expression,
and increased monoamine oxidase A expression28.
Recent collaborative eorts in amassing signicantly larger
post-mortem sample sizes, and applying more sophisticated
methods of analysis, may improve our understanding in the near
future30. However, even with these developments, the drawbacks
of post-mortem studies include heterogenous tissue quality, the
fact that the majority of samples are from older patients with a
long history of antipsychotic use, limited information regarding
clinical phenotype, and that death itself leads to a wide range
of neurobiological changes that may obscure important dier-
ences.
Studies in living participants have greater potential to include
younger individuals, drug-free subjects, and also the ability to
look at within-individual changes in symptoms and how these
relate to pharmacological manipulation.
Psychopharmacology of dopaminergic agonists and
antagonists
The discovery that chlorpromazine and reserpine were ef-
fective in the treatment of schizophrenia occurred prior to the
identication of dopamine as a neurotransmitter. It was not until
the 1970s that the clinical potency of antipsychotics was incon-
World Psychiatry 19:1 - February 2020 17
trovertibly linked to blockade of the dopamine D2 receptor31,32.
In addition, selective D2 antagonists show equivalent ecacy to
drugs with a broad spectrum of activity33, indicating that D2 an-
tagonism is sucient for antipsychotic ecacy.
It was also noted that drugs such as amphetamine that in-
crease dopaminergic neurotransmission could induce psychotic
symptoms in healthy individuals, and exacerbate psychotic
symptoms in individuals with schizophrenia34,35. Similarly, L-
DOPA treatment in Parkinson’s disease has been found to in-
duce psychotic symptoms in some individuals36. However, while
amphetamine-induced psychosis is marked by hallucinations,
delusions, paranoia, and conceptual disorganization, it is not
typically associated with negative and cognitive symptoms of the
same form as those observed in schizophrenia37. is relative
specicity to positive psychotic symptoms contrasts with gluta-
matergic models of schizophrenia (see later).
Summary of indirect findings
e ndings discussed above provide evidence that aberrant
function of the dopamine system contributes to psychotic symp-
toms (see Table1). However, these methods are unable to iden-
tify where within the brain this dysfunction is localized to and,
for the most part, cannot provide a direct link to symptoms. We
next discuss methods for in vivo imaging of the dopamine sys-
tem, which has the potential to overcome these obstacles.
Imaging dopamine in vivo
Both magnetic resonance imaging (MRI) and positron emis-
sion tomography (PET) have been used to characterize the dopa-
mine system in vivo (Table2). PET provides molecular specicity
to the dopamine system, but this comes at the cost of lower tem-
poral and spatial resolution compared to MRI.
MRI
Although MRI lacks the ability to directly image the dopa-
mine system, recent work imaging neuromelanin has shown
some promise in quantifying the dopamine system in vivo. Neu-
romelanin is synthesized via iron dependent oxidation of cyto-
solic dopamine, and accumulates in dopamine neurons of the
substantia nigra. It has been demonstrated that the neuromela-
nin MRI signal is associated with integrity of dopamine neurons,
with dopamine release capacity in the striatum, and with the se-
verity of psychosis in schizophrenia49.
Functional MRI (fMRI) has also been used in attempts to infer
functioning of the dopamine system. Task-based fMRI has been
adopted to quantify the striatal response to reward, and this has
been linked to dopamine function, although the precise relation-
ship is complex50. ere is consistent evidence of reduced ventral
striatal activation to reward in schizophrenia51. We consider how
this is consistent with the hypothesis of an overactive dopamine
system in the section discussing psychotic symptoms below.
PET: dopamine receptors
Dopamine receptors have been studied using a wide range of
radioligands. e majority of studies have used ligands specic
for D2-type (i.e., D2, D3 and D4) dopamine receptors, although
several studies have also examined D1-type (i.e., D1 and D5) re-
ceptors.
Table 1 Summary of indirect evidence for dysfunction of dopamine and glutamate systems in schizophrenia
Dopamine Glutamate
Animal models Amphetamine administration, striatal D2 overexpression, and
transgenically increased dopamine synthesis capacity are
associated with schizophrenia-like behaviours. Models of
neurodevelopmental and social risk factors are associated
with increased striatal dopamine function.
Administration of NMDA antagonists induces a wide variety
of schizophrenia-like behaviours. Genetic models that
disrupt NMDA signalling (by reducing levels of D-serine,
inactivating D-amino oxidase or decreasing dysbindin) show
behavioural and neurobiological changes similar to those
observed in schizophrenia.
Cerebrospinal fluid (CSF) Studies of DOPAC and HVA both peripherally and in CSF
have been inconsistent.
Studies of glutamate levels are inconsistent, but kynurenic acid
(an NMDA antagonist) levels appear consistently raised.
Post-mortem studies Increased D2 receptor densities have been observed, but may
result from medication use.
Glutamate neurons show reduced dendrite arborization, spine
density and synaptophysin expression. Glutamate trans-
porter EAAT2 protein and mRNA levels appear reduced
in frontal and temporal areas. There is some evidence that
glutaminase expression is increased in patients, and also that
GRIN1 abnormalities exist.
Pharmacological studies Clinical potency of antipsychotics is strongly linked to their
affinity for the D2 receptor. Amphetamines can induce
positive psychotic symptoms in healthy controls and worsen
symptoms in patients.
NMDA antagonists induce positive, negative and cognitive
psychotic symptoms in healthy controls. Chronic ketamine
users show subthreshold psychotic symptoms.
NMDA – N-methyl-D-aspartate, DOPAC – 3,4-dihydroxyphenylacetic acid, HVA – homovanillic acid, EAAT – excitatory amino acid transporter
18 World Psychiatry 19:1 - Februar y 2020
Striatum
It has been proposed that excessive dopaminergic neuro-
transmission in schizophrenia results from upregulation of stri-
atal postsynaptic D2-type receptors. However, meta-analyses of
studies using PET show only a small increase in receptor den-
sity at most in schizophrenia, and there is no signicant dier-
ence between patients and controls in analyses restricted to
medication naïve patients52. When combined with evidence that
antipsychotic treatment appears to lead to D2 receptor upregula-
tion24,25, it appears possible that any patient-control dierences
may be secondary to confounding by treatment.
ere are caveats, however, to the above inference. First, the
majority of studies are unable to measure the absolute density
of receptors, because a proportion of receptors will be occupied
by endogenous dopamine. If schizophrenia is associated with in-
creased synaptic dopamine levels, this could mask a concurrent
increase in receptor densities. Indeed, one study where dopa-
mine depletion was undertaken prior to PET scanning showed
significantly increased dopamine receptor availability in pa-
tients, although this increase was not signicant in another study
using this approach53,54.
Second, the majority of ligands are selective for D2 over D3
and D4 receptors. e studies that have employed butyrophe-
none tracers (that have an anity for D4 receptors in addition to
D2 and D3 receptors) have tended to show raised receptor den-
sities compared to those studies employing ligands that do not
have D4 anity52. In addition to potential dierences in D2/3/4
subtype proportions, D2 receptors exist in both high and low af-
nity states, and some evidence suggests that schizophrenia may
be associated with an increased proportion of receptors in the
high anity state55-58.
Furthermore, following receptor internalization, some tracers
remain bound, while others dissociate. So, if receptor internali-
zation is increased in one group, this would register as reduced
ligand binding if using a tracer that dissociates on internaliza-
tion, but not if using a tracer that remains bound59,60.
Finally, it has recently been shown that the variability of stri-
atal D2 receptor levels is greater in patients than controls61, sug-
gesting that dierences in D2 receptor density may exist, but only
within a subgroup of patients, although whether this reects a
primary pathology or an eect of prior antipsychotic treatment
in some patients remains unclear.
D1-type receptors have not been studied frequently in the
striatum, and the studies that have been undertaken do not show
any clear patient-control dierences52,62.
Extra-striatal regions
e measurement of dopamine receptors in extra-striatal re-
gions is complicated by the lower receptor densities, which means
Table 2 Summary of imaging studies of the dopamine and glutamate systems in schizophrenia
Striatal Extra-striatal
DOPAMINE
Dopamine receptors
D1 Few studies, and no differences
consistently noted.
Studies using [3H]SCH 23390 have reported
decreased binding in patients; those using [11C]
NNC 112 reported an increase in patients.
D2 No patient/control differences in
unmedicated cohorts. Variability
increased in patients.
Generally poor signal-to-noise ratio. No consistent
patient/control differences.
Presynaptic dopamine function
Consistently increased in both previously
medicated and antipsychotic naïve
patients (g=0.7). Patient-control
differences appear greatest in the
dorsal striatum.
Two studies have found increased synthesis capacity
in the substantia nigra (although not observed
in another). One amphetamine challenge study
found reduced release in patients in prefrontal
cortex. Psychological challenges have produced
less clear results. Findings of challenge studies in
the substantia nigra are inconsistent.
Dopamine transport
DAT No patient-control differences in mean
binding, but variability increased in
patients.
Fewer studies. Some suggestion that thalamic levels
may be raised in patients.
VMAT Two studies have found increased levels in the
ventral brainstem of patients.
GLUTAMATE
Basal
ganglia
Glx (g=0.4) and glutamate (g=0.6) levels raised in patients.
Thalamus Glutamine levels raised in patients (g=0.6).
Medial temporal
lobe
Glx levels raised in patients (g=0.3).
DAT – dopamine transporter, VMAT – vesicular monoamine transporter, Glx – glutamate + glutamine
World Psychiatry 19:1 - February 2020 19
that the signal-to-noise ratio is much lower than in the striatum.
Studies of thalamic, temporal cortex and substantia nigra D2/3
receptor availability have not consistently shown patient-control
dierences63. Other cortical regions have rarely been studied, and
have not shown consistent changes63.
D1 receptors have been more thoroughly examined in cortical
regions than in the striatum. Two studies using [11C]NNC 112 re-
ported an increase in patients64,65, while one reported a decrease66.
Four studies using [3H]SCH 23390 have reported a decrease62,66-68,
while two found no signicant dierences69,70. e interpretation
of these ndings is complicated by the fact that dopamine deple-
tion paradoxically decreases the binding of [3H]SCH 23390, while
it has no eect upon [11C]NNC 112 binding. Furthermore, anti-
psychotic exposure decreases D1 receptor expression, and both
the above ligands also show anity for the 5-HT2A receptor71-73.
PET: dopamine transport mechanisms
DAT is involved in reuptake of dopamine from the synaptic
cleft, and is often interpreted in PET studies as a measure of the
density of dopamine neurons. Studies examining DAT density in
the striatum have found no consistent dierences between pa-
tients and controls52, although, as with D2 receptors, variability is
increased in schizophrenia, suggesting that dierences may exist
within a subgroup61. A more recent study did nd signicantly
raised striatal DAT levels in patients, but this was observed in those
with a chronic illness with long-term antipsychotic exposure74.
ere have been fewer studies examining extra-striatal regions,
although the ones that have been undertaken do suggest that tha-
lamic DAT levels may be raised in patients74,75.
VMAT2 transports intracellular monoamines into synaptic
vesicles. Two PET studies have found that its levels were increased
in the ventral brainstem of individuals with schizophrenia, but
found no dierences compared to controls in the striatum or thal-
amus76,77. is is in contrast to the post-mortem studies discussed
above28, but in keeping with a study showing increased VMAT2
density within platelets from individuals with schizophrenia78.
PET: presynaptic dopamine function
Multiple methods exist for quantifying aspects of presynaptic
dopamine function.
Several studies have investigated dopamine release capacity
by studying the reaction of the dopamine system to an acute chal-
lenge, be that pharmacological such as amphetamine, or psycho-
logical such as a reward or stress task79. Animal studies have shown
that comparing ligand binding during the challenge to binding at
baseline allows one to infer the amount of dopamine release in-
duced by the task80.
Alternatively, one can obtain a measure of baseline synaptic
dopamine levels by comparing a baseline scan with a scan ob-
tained following the administration of a dopamine depleting agent
such as alpha-methylparatyrosine.
Finally, radiolabelled L-DOPA can be used to quantify dopa-
mine synthesis capacity. Radiolabeled L-DOPA is taken up by do-
pamine neurons, where it is converted by aromatic L-amino acid
decarboxylase to dopamine, which is then sequestered in vesicles
within nerve terminals81. e rate of uptake provides an index of
dopamine synthesis capacity.
Striatum
Studies have consistently demonstrated raised presynaptic do-
pamine function in schizophrenia, with Hedges’ g=0.7 (Hedges’ g
is a measure of eect size, and values of 0.2 are typically considered
small, those of 0.5 medium, and those of 0.8 large82). e studies
using a challenge paradigm show larger eect sizes (g=1.0) com-
pared to those quantifying dopamine synthesis capacity (g=0.5)83.
e hyperdopaminergic state associated with schizophrenia ap-
pears greatest within the dorsal striatum83.
Further evidence for pathophysiological relevance comes from
studies showing a direct association between synthesis capac-
ity and the severity of positive psychotic symptoms84-86. e re-
lationship with other symptom domains is less clear: an inverse
relationship with depressive symptoms87 and a lower synthesis ca-
pacity associated with worse cognitive performance88 have been
reported.
Extra-striatal regions
Outside of the striatum, dopamine synthesis capacity can only
be reliably measured in a limited number of brain regions, such as
the substantia nigra and the amygdala, using current techniques89.
Two studies have found increased dopamine synthesis capac-
ity in the substantia nigra90,91, although this was not observed in
another92. One study also found raised dopamine turnover in the
amygdala91.
Although cortical dopamine receptors are predominantly D1-
type, D1 receptor ligands are not reliably displaceable, and there-
fore not suitable for challenge or displacement studies. Cortical D2
receptors do exist, but studies are complicated by their sparsity93.
Furthermore, although challenge paradigms have demonstrated
validity in the striatum, the results of cortical studies are harder to
interpret, with displacement not always observed94.
One study using amphetamine challenge in combination with
the high-anity ligand FLB-457 found reduced dopamine release
in the prefrontal cortex in individuals with schizophrenia95. Two
other recent FLB-457 studies adopted psychological challenges.
One of these used a psychological stressor, which did not induce
cortical tracer displacement in either patients or controls96. e
other used a cognitive test of executive function, which did show
lower tracer displacement in patients, but interpretation was com-
plicated by the fact that, again, the task did not consistently induce
dopamine release97. A study using 18F-fallypride found no dier-
ences between patients and controls in terms of stress-induced
cortical dopamine release98.
20 World Psychiatry 19:1 - Februar y 2020
Two studies have examined dopamine release in the substan-
tia nigra. One used a stress challenge and found an increased re-
lease in patients99; the other adopted an amphetamine challenge
and found a non-signicant reduction95.
PET: dopamine across the psychosis spectrum
Several studies have investigated dopamine function in sub-
jects at clinical high risk for psychosis. Initial studies showed
evidence of raised presynaptic dopamine function in these in-
dividuals100-102. However, this was not seen in the largest study
to date103. is may potentially result from the fact that raised
presynaptic striatal dopamine function appears to be limited to
those subjects who subsequently develop psychosis104, and tran-
sition rates have declined in recent years.
A study of healthy individuals that experience auditory hallu-
cinations also found no dierence in striatal dopamine synthesis
capacity compared to healthy controls without hallucinations105.
Studies in individuals at increased genetic risk for schizophrenia,
such as patients’ relatives and individuals with 22q11.2 deletion
syndrome, have also not shown consistent dierences from con-
trols in terms of presynaptic dopamine function106-108.
Studies in psychotic individuals with diagnoses other than
schizophrenia, such as bipolar disorder and temporal lobe epi-
lepsy86,109, have found raised striatal dopamine synthesis capac-
ity. is nding, along with the inconsistent evidence in people
at increased clinical or genetic risk, may suggest that increased
striatal dopamine synthesis capacity is associated with psycho-
sis across psychiatric diagnoses, rather than being an underlying
risk factor for schizophrenia.
Studies of dopamine receptor densities in individuals at both
clinical99,100,110 and genetic107,110-113 high risk are similar to those
in individuals with schizophrenia, in that they have shown no
clear dierences from controls.
Summary of PET findings
e studies reviewed above provide consistent evidence of a
striatal presynaptic hyperdopaminergic state in schizophrenia
(see Table2), and little consistent evidence of altered D2/3 re-
ceptor levels. It remains uncertain as to whether abnormalities
exist with regard to other dopamine receptors, or with cortical
dopamine function.
Consequences of dopaminergic dysfunction
Prediction errors, salience and positive symptoms
After its role in movement was established, preclinical nd-
ings suggested that dopamine also played a role in signalling
reward114. Later work demonstrated that signalling more speci-
cally related to the discrepancy between expected and received
reward – a reward prediction error115. More recently, it has been
demonstrated that ring is not exclusively tied to reward predic-
tion, but rather can occur in response to a wide range of salient
stimuli116-120, and that in more dorsal regions of the striatum do-
pamine signalling is particularly associated with threat-related
stimuli118,119.
Several related theories have proposed how disruption to nor-
mal dopamine function could underlie positive psychotic symp-
toms such as delusions and hallucinations121-123. Dysregulated
dopamine neuron ring will aberrantly signal that irrelevant stim-
uli are of importance, thereby imbuing percepts and thoughts with
abnormal salience, in turn leading to inappropriate associations
and causal attributions124. ere are also mechanisms through
which uncoordinated dopamine signalling may contribute not
only to the generation of delusional beliefs, but also to the impervi-
ously rigid form of delusional thought123,125.
Recent work has attempted to identify more precisely the
mechanisms through which dopaminergic dysfunction may
contribute to symptoms. e experience of a stimulus depends
not only on the sensory inputs resulting from that stimulus, but
also on prior expectations regarding the probability of a percept.
Auditory hallucinations appear to result from a stronger inu-
ence of prior expectations upon sensory percepts126, and this
increased weighting of priors has been associated with greater
levels of amphetamine-induced dopamine release in the stria-
tum127.
In terms of understanding the development of delusions, a
combined PET and MRI experiment found that dopamine re-
lease was related to neural signalling of belief updates rather
than just sensory surprise128. is suggests that aberrant dopa-
mine signalling may lead to irrelevant stimuli being understood
as meaningful, the clinical relevance of which is supported by
the nding that participants who displayed more aberrant belief
updating showed greater subclinical paranoid ideation128.
In addition to mesostriatal dopamine signalling, several cor-
tical regions have also been implicated in salience process-
ing129,130. e salience network comprises the anterior cingulate
cortex and bilateral insula, and abnormalities of this network
have been proposed as a core feature of schizophrenia patho-
physiology131. e network has a key role in orchestrating dy-
namic switching between brain states, for example between a
resting state and states associated with performing cognitively
demanding tasks132. It is of relevance that dopamine signalling
also plays a role in dynamic reorganization of brain states133,134.
Recent work has demonstrated that mesostriatal dopamine sig-
nalling and salience network connectivity are tightly linked135,
although whether this relationship is disrupted in schizophrenia
is not known.
Reward, motivation and negative symptoms
Reward and punishment are fundamental drivers of behav-
iour, and reinforcement learning models formalize the relation-
ship between reward, states and behaviour. Prediction errors
World Psychiatry 19:1 - February 2020 21
allow the value of states and actions to be learnt, and are a key
signal in many reinforcement learning models. Given the central
role of dopamine in both coding prediction errors and in the cor-
tical representation of environmental states, several studies have
used this framework to explore the behavioural consequences of
disrupted dopamine signalling136.
Cortical D1 receptors play a central role in shaping the accu-
rate neural representation of environmental states, by allowing
precise inhibition of neural activity137. Reduced cortical dopa-
mine signalling means that stimuli associated with reward can-
not be accurately encoded, eectively foreclosing their ability to
guide behaviour137. Furthermore, reduced cortical dopamine
signalling may mean that reward-related representations are
short-lived, with the consequence that, even if correctly repre-
sented, the motivational properties of reward-associated stimuli
have a briefer impact137.
Dopamine neurons re in response to stimuli that have been
previously associated with reward, and guide behaviour towards
actions associated with previous reward138. A striatal hyperdo-
paminergic state may mean that reward-associated stimuli have
reduced motivational inuence, as aberrantly high background
levels of dopamine signalling reduce the signal-to-noise ratio
of adaptive phasic signalling139. is mechanism also has the
potential to reduce the appetitive properties of a given reward,
thereby reducing its impact to shape future behaviour, and ac-
counting for negative symptoms such as anhedonia and amoti-
vation140-142. is reduced signal-to-noise ratio may account for
the reduced striatal activation to reward observed with fMRI in
individuals with schizophrenia51.
Cortical dopamine and cognitive symptoms
Cognitive symptoms of schizophrenia include deficits in
working memory, executive function, and information process-
ing. ey occur prior to the onset of frank psychosis and account
for a signicant proportion of the morbidity associated with the
illness143-145.
e dorsolateral prefrontal cortex is central to many cognitive
processes, and both functional and structural pathology of the
region has been linked to the decits seen in schizophrenia146.
e molecular changes underlying cognitive symptoms, howev-
er, are unknown. Given the importance of D1 receptor signalling
for cognition147, reduced cortical dopamine signalling has long
been hypothesized to contribute to the cognitive symptoms ob-
served in schizophrenia. As discussed above, however, evidence
regarding D1 receptors in schizophrenia is inconclusive, and
there has been only a single study demonstrating reduced corti-
cal dopamine release95.
On the basis of preclinical work using a model of striatal D2
overexpression, it has also been proposed that excessive striatal
dopamine signalling may lead to reductions in cortical dopa-
mine and associated cognitive symptoms7. erefore, it appears
that both reduced and excessive dopamine signalling can have
deleterious eects on cognition, which may contribute to the fact
that there has been minimal success in developing dopamine
modulating treatments for cognitive symptoms of schizophre-
nia4,145.
GLUTAMATE
As reviewed above, there is signicant evidence that dysfunc-
tion of the dopamine system is involved in the pathogenesis of
schizophrenia. is may, however, occur downstream of other
pathophysiological processes. Furthermore, schizophrenia is a
heterogenous disorder, and dopaminergic dysfunction may not
play a signicant role in some individuals, while in others it may
be only one of several pathological mechanisms.
Substantial evidence has accumulated implicating the gluta-
mate system in the pathogenesis of schizophrenia. Glutamate
is the major excitatory neurotransmitter in the central nervous
system, and, in contrast to the anatomically localized cell bod-
ies of dopamine neurons, glutamatergic neurons are widespread
throughout the brain.
Glutamate receptors show considerable variety, and are clas-
sied as either ionotropic or metabotropic. Ionotropic receptors
include the NMDA and the non-NMDA receptors – kainate and
alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid
(AMPA). NMDA and non-NMDA receptors are normally co-lo-
calized on neurons, and act synergistically in that the NMDA re-
ceptor has slower kinetics and tends to enhance depolarization
initiated by non-NMDA receptors.
Following the discovery of the psychotomimetic effects of
NMDA antagonists, the NMDA receptor has become a primary
focus when considering potential glutamatergic dysfunction
in schizophrenia. Below we examine the evidence associating
schizophrenia with glutamatergic abnormalities, consider how
these abnormalities may lead to symptoms, and review the po-
tential of glutamate modulating agents as treatments.
Indirect evidence for glutamatergic dysfunction in
schizophrenia
Animal models
The administration of NMDA antagonists to non-human
primates and rodents has been shown to induce a variety of
schizophrenia-like behaviours, such as sensorimotor gating im-
pairments, increased locomotion, abnormal repetitive move-
ments, and cognitive and social decits38. NMDA antagonism
has also been shown to lead to hippocampal hypermetabolism,
similar to that which has been observed in schizophrenia148,149.
A genetic animal model that reduced levels of the NMDA co-
agonist D-serine was associated with neurobiological changes
similar to those observed in schizophrenia, such as reduced
dendritic spine density and hippocampal volume39. Several
other genetic mice models, such as those involving inactivation
of D-amino oxidase40 and reduction of the synaptic protein dys-
22 World Psychiatry 19:1 - Februar y 2020
bindin41, also show NMDA receptor hypofunction accompanied
by behavioural and neurobiological changes analogous to those
observed in schizophrenia.
CSF and post-mortem studies
Initial small studies of CSF found reduced glutamate levels
in patients, but these ndings were not replicated42,43. However,
CSF and post-mortem brain levels of kynurenic acid, an NMDA
receptor antagonist, have consistently found to be raised, al-
though not plasma levels44.
Post-mortem studies investigating structural alterations of
glutamate neurons have generally found reductions in den-
drite arborization, spine density, and synaptophysin expression
across frontal and temporal regions45.
mRNA expression of specic glutamate receptors and their
subunits has also been investigated across multiple brain re-
gions. Although several studies have found reduced expres-
sion of NMDA receptor subunits such as GRIN1, GRIN2A and
GRIN2C, these have not been consistently replicated across
brain regions45, although – when regions are examined individu-
ally – there is evidence of reduced GRIN1 expression in the hip-
pocampus45. A recent study examining samples from over 500
individuals with schizophrenia found increased exon skipping in
GRIN1, that would aect the extracellular ligand binding site30.
Fewer studies have examined protein expression of these subu-
nits, and these are also inconsistent in their ndings45.
ere have also been a small number of studies investigating
enzymes involved in glutamate metabolism. Excitatory amino
acid transporters (EAAT) remove glutamate from the synaptic
cleft. After reuptake from the synapse, glutamine synthetase con-
verts glutamate to glutamine. When glutamine is delivered to neu-
rons, it is converted back to glutamate by glutaminase. ere is
some preliminary evidence that both mRNA and protein levels of
EAAT2, the transporter responsible for the majority of glutamate
uptake, are reduced in frontal and temporal areas of patients with
schizophrenia, while glutaminase mRNA levels and enzymatic
activity have been found to be raised in two studies examining,
respectively, the thalamus and dorsolateral prefrontal cortex45.
Although the caveats regarding post-mortem studies dis-
cussed above remain, several of the ndings covered here are
consistent with impaired functioning of the NMDA receptor.
In addition to the direct pathology suggested by the findings
relating to GRIN1, the reduced expression of EAAT2 and the
increased expression of glutaminase may be understood as com-
pensatory responses attempting to increase synaptic glutamate
levels.
Psychopharmacology of NMDA antagonists
e administration of NMDA receptor antagonists, such as
phencyclidine and ketamine, to healthy controls has been re-
peatedly shown to induce manifestations similar to the posi-
tive, negative and cognitive symptoms of schizophrenia46. It
has been demonstrated that NMDA receptor blockade by these
compounds is both necessary and sucient for their psychoto-
mimetic eects150.
ese drugs have also been shown to exacerbate a similarly
wide spectrum of symptoms in individuals with schizophrenia,
in contrast to amphetamines, which predominantly worsen
positive symptoms151. Furthermore, it has been shown that anti-
NMDA receptor encephalitis may be associated with psychiatric
presentations that resemble schizophrenia in some individu-
als152. NMDA antagonism has, therefore, been proposed to be a
superior pharmacological model of schizophrenia, compared to
amphetamines, given its ability to more reliably induce negative
as well as positive symptoms153,154.
Acute NMDA antagonist administration has typically been
employed in experimental settings, since chronic administration
to healthy controls cannot be ethically undertaken. However,
chronic ketamine users show subthreshold psychotic symptoms
across positive, negative and cognitive domains47,48, and a sub-
group of these chronic users develop a persisting psychosis that
is more similar to schizophrenia than that induced by acute sin-
gle dose ketamine administration155.
Imaging glutamate in vivo
Proton magnetic resonance spectroscopy
Proton magnetic resonance spectroscopy (1H-MRS) is the
most frequently used technique for investigating the glutamate
system in vivo. It does not require the use of ionizing radiation,
and is signicantly cheaper than PET.
1H-MRS does, however, have several drawbacks, including the
fact that it is unable to distinguish between intra- and extracel-
lular compartments156. Glutamate does not act solely as a neu-
rotransmitter, but is involved in protein synthesis and nitrogen
metabolism, and is a precursor to GABA157. is means that it is
not possible to infer whether dierences between patients and
controls relate to synaptic glutamate concentrations, as opposed
to alterations in these other functions of glutamate. Even if one as-
sumes that detectable abnormalities relate to dierences in syn-
aptic neurotransmission, it is not possible to determine whether
this is secondary to dierences in synaptic levels, as opposed to
presynaptic dysfunction, or altered reuptake of glutamate.
At higher eld strengths, glutamate and its metabolite glu-
tamine can be distinguished, while at lower strengths the con-
centration of both, often abbreviated to Glx, is all that can be
accurately quantied. Glutamine is synthesized from glutamate
following the uptake of synaptic glutamate by astrocytes, and
glutamine levels have been taken to be a marker of glutamate
neurotransmission. However, as with glutamate, glutamine takes
part in multiple cell process, which complicates interpretation158.
ere have been over fty 1H-MRS studies of the glutamate
system in schizophrenia. A synthesis of their ndings is com-
plicated by methodological dierences, notably in the imaging
World Psychiatry 19:1 - February 2020 23
sequences used, the brain regions studied, and the patient co-
horts enrolled. Notwithstanding these issues, a meta-analysis of
these studies has found several relatively consistent ndings159.
ere is evidence that Glx (g=0.4) and glutamate (g=0.6) levels
are raised in the basal ganglia, that glutamine concentration is
increased in the thalamus (g=0.6), and that Glx levels are raised
in the medial temporal lobe (g=0.3) in patients with schizophre-
nia (Table2).
Although these eect sizes are in some instances comparable
to those observed for presynaptic dopamine function, they rep-
resent considerably less studies. In the case of both glutamate in
the basal ganglia and glutamine in the thalamus, only three stud-
ies have been performed, and as such these ndings should be
regarded as preliminary. e increased Glx levels observed in the
medial temporal lobe represent 18 studies (13 in schizophrenia
and 5 in clinical high-risk individuals).
If taken to reect synaptic glutamate levels, the temporal lobe
ndings are consistent with the post-mortem ndings discussed
above, and the PET study discussed below, which suggest that
receptor dysfunction is accompanied by compensatory changes
that increase synaptic glutamate levels. is is also consistent
with studies that have found hippocampal hyperactivity in psy-
chosis, potentially resulting from dysfunction of NMDA recep-
tors located on GABAergic interneurons9,148,149.
In recent years, several studies have been performed using high-
er magnetic eld strengths of up to 7 Tesla. Higher eld strength
has greater sensitivity, enabling greater separation of glutamate
and glutamine peaks, and allowing for more accurate quantica-
tion. Two studies in rst-episode patients have shown reduced glu-
tamate concentrations in the anterior cingulate cortex160,161, while
another only found this in a subset of patients with predominantly
negative symptoms162. No dierence has been observed in several
other investigations163-166. Overall this suggests that, even at high
eld strengths, group dierences are inconsistent.
Another recent development is functional MRS. is involves
acquiring multiple measures during a task or other stimulus, to
investigate dynamic changes in metabolite measures. Changes
in 1H-MRS measures during a stimulus are proposed to result
from compartmental shifts, as glutamate located extracellularly
may contribute more to the signal. is results from the fact that,
within presynaptic vesicles, metabolite movement is restricted,
and therefore may have a faster T2 relaxation rate167.
A study using a heat pain stress found a reduced anterior cin-
gulate cortex glutamate response in individuals with schizophre-
nia compared to healthy controls168, although interpretation is
complicated by the fact that baseline glutamate levels were high-
er in patients. A similar pattern was seen in a study using a cog-
nitive task in which patients showed reduced anterior cingulate
glutamate response compared to controls166.
Other neuroimaging techniques
Given the limitations of 1H-MRS in determining the precise
nature of glutamatergic abnormalities, several attempts have
been made to develop radioligands capable of directly measur-
ing glutamate receptors.
One study found reduced NMDA receptor binding in the left
hippocampus of patients with schizophrenia, but no further
studies have been attempted, in part due to concerns regarding
a lack of specicity of the tracer. A recent study using a tracer for
the metabotropic glutamate receptor 5 found no patient-control
dierences169. New tracers are under development, but lack of
specicity remains an ongoing problem when attempting to im-
age glutamate receptors170.
13C-MRS is another non-invasive imaging technique. It has
lower sensitivity compared to 1H-MRS, but, when combined with
a 13C labelled infusion, it has the potential to overcome some of
the limitations associated with 1H-MRS, specically as regards to
characterizing the glutamate-glutamine cycle171. is technique
has been adopted to show that the majority of energy production
in the brain supports glutamatergic activity and, although stud-
ies are yet to be performed in schizophrenia, it has recently been
used in humans to show that ketamine increases glutamate-glu-
tamine cycling172,173.
Glutamate chemical exchange saturation transfer (GluCEST)
is another novel technique for measuring glutamate in vivo. In
addition to improved sensitivity, it allows for whole brain imag-
ing without the need to specify a single voxel174. It has so far been
used in a single investigation, which reported reduced glutamate
levels in individuals with schizophrenia and those at clinical
high risk, compared to healthy controls175.
Glutamate across the psychosis spectrum
As with imaging of the dopamine system, eorts have been
made to characterize glutamatergic function in individuals at clin-
ical and genetic high risk for psychosis. Similarly, although a me-
ta-analysis suggested that increased Glx levels might exist in the
medial frontal cortex in individuals at clinical high risk of psycho-
sis165, recent studies have been negative176-178, and there appear to
be no clear dierences between at-risk individuals and controls.
Given that 1H-MRS measures of glutamate have been consist-
ently shown to decline with age across the frontal cortex, anterior
cingulate, hippocampus and basal ganglia179-181, it may be that
schizophrenia is associated with a distinct trajectory of change, and
so dierences only become detectable later in the time course of
the illness.
Consequences of glutamatergic dysfunction
Unlike dopamine neurons, which are restricted to relatively
well circumscribed anatomical pathways, glutamate signalling
occurs ubiquitously throughout the brain and, as a result, dys-
function of this system has the potential to account for a wide
range of impairments.
However, given the limitations regarding techniques for direct-
ly quantifying the glutamate system in vivo, there is a paucity of
24 World Psychiatry 19:1 - Februar y 2020
direct evidence regarding the precise nature of glutamatergic
dysfunction in schizophrenia, and studies looking at the rela-
tionship between 1H-MRS measures of glutamate and symptom
severity have produced inconsistent ndings182. Indeed, both
increased and decreased level of glutamate as measured by 1H-
MRS have been proposed to support a hypothesis of NMDA hy-
pofunction in schizophrenia183,184.
NMDA receptors, sparse coding and memory
NMDA receptors play a vital role in orchestrating several cog-
nitive processes, including working memory185. One of the mech-
anisms involved in the ecient cortical representation of in-
formation is that of sparse coding.
Dierent cortical cell classes encode information dierently.
More supercial cell layers code “sparsely”. is means that only
a small proportion of cells within a region will be spiking, and in-
formation is thus encoded spatially186. is is in contrast to deeper
layers, where the majority of cells may be ring, and information
is encoded by variations in the rate of spiking186. Sparse coding in-
volves great spatial precision in terms of the area of excitation, and
this is mediated by strong lateral inhibition secondary to dense
GABAergic interneuron networks within the supercial layers187.
Sparse coding allows the maintenance of multiple mnemonic
networks, and also the protection of encoded memories from
distractors137. Disruption to sparse coding mechanisms has been
shown to lead to the development of false memories in animal
models188. Both individuals with schizophrenia and healthy
controls administered NMDA antagonists display phenomena
that may result from impaired sparse coding – a smaller work-
ing memory buer, decreased mnemonic precision, and false
alarms in working memory137,189. Decreased inhibition second-
ary to hypofunction of NMDA receptors on GABAergic interneu-
rons could account for a disruption to the spatial precision of
supercial layer ring, but this remains to be denitively tested.
Excitatory/inhibitory balance and neuronal oscillations
Synchronized neuronal oscillations are associated with a wide
range of cognitive processes, such as working memory. ese
oscillations result from a tightly maintained balance between
excitatory and inhibitory populations of neurons, and can be
measured in vivo using electroencephalogram (EEG).
e balance between excitation and inhibition is crucial for
normal physiological function, and NMDA receptors play a criti-
cal role here. Disruption to this balance has been proposed to
result in the EEG abnormalities observed in schizophrenia. is
disorder is associated with increased resting gamma oscillations,
that have been linked to cognitive symptoms190. is disruption
of normal oscillatory activity mirrors what has been observed
with ketamine administration in healthy volunteers.
A range of molecular alterations have been proposed to po-
tentially lead to excitatory/inhibitory imbalance in schizophre-
nia, including excessive pruning of dendritic spines, and intrinsic
GABAergic abnormalities191. Another candidate mechanism is
that of NMDA hypofunction. NMDA antagonists preferentially
act on GABA interneurons, because these neurons have a more
depolarized membrane potential. It has also been proposed that,
in schizophrenia, NMDA hypofunction may preferentially aect
interneurons192, which would in turn lead to greater activity of
pyramidal neurons. is uncoordinated increased activity may
underlie disruptions to normal oscillatory activity mentioned
above, and act as noise, impairing the ability of coordinated ac-
tivity to be passed down to subcortical regions150.
FACTORS UNDERLYING GLUTAMATERGIC AND
DOPAMINERGIC DYSFUNCTION
Genetic factors
Over a hundred risk loci have been associated with schizophre-
nia by large scale genome-wide association studies (GWAS)195.
Given the evidence implicating dopamine in the disorder, it is sur-
prising that only one of the loci was found to be associated with
the dopamine system, specically the dopamine D2 receptor.
Analyses specically looking at genes involved in dopamine
synthesis, signalling and metabolism have also been unable to nd
a signal other than for D2 receptors, and this accounts for a negli-
gible proportion of the overall genetic risk196. However, other loci
strongly linked to schizophrenia, such as 10q24.32, have shown
associations with dopamine synthesis capacity197, as have poly-
morphisms of the gene DISC1198. Although relatively small sample
sizes were used, these ndings suggest that genetic factors related
to schizophrenia may indirectly aect the dopamine system.
In the case of glutamate, many genes involved in the develop-
ment and maintenance of glutamatergic synapses are not only
implicated by loci signicantly associated with schizophrenia in
GWAS193,195, but also by studies examining rarer genetic risk fac-
tors199,200. is includes genes that directly code for components
of glutamate receptors, such as GRIN2A, GRIA1 and GRM3, and
genes involved in facilitating glutamatergic neurotransmission
through other means, such as that coding for serine racemase
(SRR). ese results provide some of the strongest support that
disruption of glutamatergic signalling is a fundamental compo-
nent of schizophrenia pathophysiology.
Induced pluripotent stem cells (iPSCs) allow for the genera-
tion of live neurons, in vitro, from somatic cells taken from pa-
tients. ese neurons are best conceptualized as modelling fetal
brain tissue, primarily reecting an underlying genetic architec-
ture, distilled from environmental exposures201. e use of iPSCs
can be particularly valuable in elucidating the eects of genetics
in a polygenic disorder such as schizophrenia, where disease risk
is encoded by complex networks of genes, the functional conse-
quences of which are not easily intuited.
Several studies using this technique have investigated how
the dopamine system may be aected. One study found reduced
DAT expression, indicating immaturity in dopaminergic neurons
derived from individuals with schizophrenia202. Another study,
however, found that dopamine neurons derived from patients
World Psychiatry 19:1 - February 2020 25
tended to develop more rapidly, and showed increased dopa-
mine release203. Small sample sizes and dierences in experi-
mental protocols likely contributed to these discrepancies204.
iPSCs have also been used to study the glutamate system in
neurons derived from individuals with schizophrenia. Two stud-
ies have demonstrated decits in glutamate receptor signalling in
patient-derived cells205,206, and a follow-up study identied spe-
cic gene modules associated with these decits in glutamatergic
signalling201. One investigation found reduced glutamate release
in schizophrenia-derived cells dierentiated to hippocampal den-
tate gyrus cells207, while another found delayed maturation of both
dopaminergic and glutamatergic neurons, and that glutamatergic
neurons displayed a reduced ability to form synaptic contacts202.
Environmental factors
Acute psychosocial stress has been shown to induce striatal
dopamine release208,209. Measures of presynaptic dopamine func -
tion are raised in migrants, and those that have experienced
childhood trauma, both of which are risk factors associated with
schizophrenia210-213. However, another risk factor for schizo-
phrenia, heavy cannabis use, is associated with blunted dopa-
mine synthesis capacity and release214,215.
Although acute stress has been shown to increase cortical glu-
tamate level in preclinical models, this has not been demonstrat-
ed in humans using 1H-MRS216,217. 1H-MRS studies in cannabis
users generally report reduced glutamate levels in both cortical
and subcortical areas, which is in keeping with animal work218.
Of note, a recent iPSC study showed that, in neurons derived
from individuals with schizophrenia, tetrahydrocannabinol ad-
ministration led to depressed glutamate signalling219.
Neural circuits and dopamine-glutamate interactions
e evidence discussed above suggests that, while the dopa-
mine hypothesis can account for the positive symptoms of psy-
chosis, it is less clear whether it can fully account for negative and
cognitive symptoms. Similarly, while glutamatergic models of
psychosis are able to replicate a wide range of symptoms of psy-
chosis, they do not directly account for the nding of increased pre-
synaptic striatal dopamine function, nor the clinical eectiveness
of dopamine antagonists. is suggests that dysfunction in both
systems contributes to the pathophysiology of schizophrenia, and
highlights the need to understand how these two systems may in-
teract.
Much research has investigated dopamine-glutamate rela-
tionships in humans using pharmacological challenges. Am-
phetamine administration has been shown to increase cortical
glutamate levels, as measured using 1H-MRS220, but dopamine
antagonists do not have consistent eects on glutamate levels as
measured using 1H-MRS221. Several, but not all, PET studies have
found that ketamine administration is associated with striatal
dopamine release222. While glutamatergic dysfunction mayen-
courage dopaminergic disinhibition, it is clear that this is not
theonlyroute to symptoms, given that dopamine antagonists do
not entirely ameliorate the eects of NMDA antagonists223.
Recent studies have combined PET and MRI measurements
in the same individuals to investigate this relationship without
the use of pharmacological modulation. One study in healthy
individuals found that increased dopamine synthesis capacity
in the ventral striatum was associated with both reduced corti-
cal and increased striatal levels of glutamate224. is is in keep-
ing with the imaging studies discussed above, in which both
increased dopamine and glutamate measures were observed
in the striatum in schizophrenia, which could potentially result
from increased activity of glutamatergic projection to the stria-
tum. Other studies found the same relationship between cortical
glutamate and striatal dopamine in clinical high-risk and rst-
episode psychosis patients, but not in controls225,226.
eories linking glutamate and dopamine have proposed that
defective NMDA receptors on cortical GABA interneurons result
in inadequate inhibition of glutamatergic projections to the mid-
brain. is, in turn, may overstimulate mesostriatal dopamine
neurons. In order to account for purported cortical dopamine
decits, it has also been proposed that overactive glutamatergic
projections might overstimulate GABA interneurons in the ventral
tegmental area, and thereby overinhibit mesocortical projection
neurons227.
e rst of these proposed mechanisms has support from
studies demonstrating ketamine-induced release of striatal
dopamine. It is also in line with the 1H-MRS studies discussed
above, if the increased basal ganglia glutamate levels observed in
schizophrenia are taken to represent increased activity of cortical
projections. e second mechanism remains speculative222.
It is likely, however, that the relationship between glutamate
anddopamine is not one-way but bidirectional. As mentioneda-
bove, human studies have suggested that modulation of the dopa-
mine system aects cortical glutamate levels. Preclinical studies
have shown that mice genetically modied to have upregulated
dopaminergic signalling display disrupted glutamatergic signal-
ling of thalamocortical neurons228. Furthermore, cortical dopamine
receptors have been shown to inuence local glutamate release229.
e wide range of pathways potentially linking the two sys-
tems, and the potential for opposing eects depending on the
number of interneurons within a circuit, means that it is not pos-
sible to disentangle precise mechanisms with currently available
neuroimaging methods.
TREATMENT
Dopamine modulating treatments
Antipsychotics are eective treatments for positive symptoms
in the majority of patients with a diagnosis of schizophrenia.
However, about one third of patients show persistent positive
symptoms despite treatment, and this has been termed treat-
ment-resistant schizophrenia230.
26 World Psychiatry 19:1 - Februar y 2020
Recent studies have suggested that dopamine synthesis ca-
pacity may predict which patients respond to treatment. An
initial study found that dopamine synthesis capacity was only
raised in those with a treatment-responsive illness84, which is
consistent with a more recent prospective study231.
Even when eective in treating positive symptoms, dopamine
antagonists do not typically show signicant benet for negative
and cognitive symptoms. is is expected if these symptoms do
not primarily relate to the hyperdopaminergic state, and particu-
larly so if they result from decits in dopaminergic signalling.
All currently licensed antipsychotics exert their dopaminergic
eects primarily at postsynaptic D2 receptors, which is down-
stream of the presynaptic hyperdopaminergic state that has
been observed in molecular imaging studies. ere are currently
a number of treatments in development which attempt to correct
dysregulated dopamine function further upstream.
Apomorphine was shown to be an ecacious treatment in an
early clinical trial232. Although later studies did not consistently
replicate this nding4, a recent investigation suggested that this
drug may normalize dopaminergic activity, potentially via ago-
nism of presynaptic autoreceptors233. Another upstream approach
involves agonism of trace amine type 1 receptors. is has been
shown preclinically to both reduce midbrain dopamine neuron
activity, and reduce the locomotor response to amphetamine234.
Finally, the PET studies discussed above have demonstrated
that presynaptic dopaminergic dysfunction is greatest in thedor-
sal striatum83, and muscarinic receptor 4 positive allosteric mod-
ulators specically inhibit dorsal striatal dopamine release, with
ecacy demonstrated in some clinical trials235,236.
e dopamine system may also be more indirectly regulated
via upstream circuits. e potential of glutamate signalling to
modulate dopamine neurotransmission has been discussed a-
bove.Reduced functioning of GABAergic interneurons has also
been suggested to contribute to disinhibition of dopamine neu-
rons, and alpha 5 selective GABA agonists have been proposed
as means of addressing this. Although neuroimaging and pre-
clinical work provides conceptual support, ecacy in patients is
yet to be demonstrated4.
ere is also the chance of intervening downstream of post-
synaptic receptors. Stimulation of D2 receptors inhibits cyclic
adenosine monophosphate (cAMP) production, while phospho-
diesterase inhibitors have an opposing eect by preventing cAMP
breakdown. Phosphodiesterase inhibitors may therefore have the
potential to block the downstream eects of excessive D2 signal-
ling, and also the potential benet of boosting cortical D1 signal-
ling237. Although clinical trials testing these compounds have not
yet been successful4, there is a signicant variability in the region-
al expression of phosphodiesterase inhibitor subtypes, and those
showing the greatest cortical expression remain to be tested238.
Glutamate modulating treatments
Glutamate modulating treatments for schizophrenia fall into
two camps: those that aim to augment NMDA signalling, and
those that aim to reduce levels of synaptic glutamate proposed
to be pathologically raised as a result of NMDA hypofunction.
A general challenge for the development of glutamatergic treat-
ments is that they will typically have relatively global effects,
whereas pathology may be conned to discrete cell types such as
NMDA receptors on specic GABAergic interneurons239,240.
Since directly augmenting synaptic glutamate levels could
have pathologically excitotoxic effects, efforts at augmenting
NMDA signalling have focused on the receptor’s glycine modu-
latory site. For activation of the NMDA receptor, glycine or D-ser-
ine has to bind to the glycine modulatory site on GluN1 subunit,
in addition to glutamate binding at the GluN2 subunit. Agonists
of the glycine modulatory site – including glycine, D-serine and
D-cycloserine – have demonstrated the ability to attenuate the
psychotogenic eects of NMDA antagonists in preclinical stud-
ies, although this has not been clearly demonstrated in humans
150.
A meta-analysis of clinical trials in individuals with schizophre-
nia suggested that D-serine may be eective in the treatment
of negative symptoms241, but a relatively large trial was subse-
quently negative242. e same meta-analysis also suggested that
glycine may have a benet for overall symptoms as well, but large
scale trials are needed for clear conrmation of this eect241. A
meta-analysis of glutamate positive modulators in the treatment
of cognitive symptoms, including D-serine and glycine, did not
demonstrate benet over placebo243.
Poor blood-brain barrier penetration by glycine and some of
the other co-agonists means that occupancy of the glycine mod-
ulatory site may be insucient to exert clinically measurable
eects. To address this, an alternative approach has involved at-
tempts to increase synaptic glycine levels by blocking the glycine
type 1 transporter, thereby inhibiting removal of glycine from the
synapse.
Bitopertin, a glycine type I transporter inhibitor, appeared to
be a promising compound in this regard, showing good blood-
brain barrier penetration, with encouraging results in early clini-
cal trials244. However, later trials were not successful, perhaps
due to the unusually high placebo responses, or the use of chron-
ic patient populations, given that there is some evidence that in-
tervention is likely to be more eective in early illness stages245.
More recently, another glycine type I transporter inhibitor was
shown to increase long-term potentiation (a marker of neuro-
plasticity) in individuals with schizophrenia, and the compound
awaits testing in clinical trials246.
Another potential approach may be to lower levels of kynuren-
ic acid, which is an endogenous antagonist of the glycine modu-
latory site247. Cyclooxygenase-2 inhibitors reduce kynurenic acid
levels, and celecoxib has shown benet as an adjunctive treat-
ment in early psychosis, but not chronic illness248. However, a
recent in vitro study found that celecoxib did not signicantly
reduce kynurenic acid levels, while parecoxib and niumic acid
did249. So, it may be that other cyclooxygenase-2 inhibitors have
the potential for greater ecacy.
Based on ndings that NMDA antagonism increases synaptic
glutamate levels, the second general approach has focused on
World Psychiatry 19:1 - February 2020 27
hypotheses that NMDA hypofunction may result in pathological-
ly raised levels of glutamate, and therefore inhibiting the release
of glutamate from terminals may be therapeutic150,250.
e metabotropic glutamate 2 receptor (mGluR2) is located
presynaptically on glutamate neurons, where it acts as an autore-
ceptor to regulate glutamate release251. Positive allosteric modu-
lators of mGluR2 have been found to be eective in reducing the
cognitive impairments induced by ketamine252. However, eca-
cy in clinical trials has not been consistently shown253. One com-
plication in these trials was the high rate of placebo response.
Riluzole (2-amino-6-triuormethoxy benzothiazole) has also
been shown to reduce synaptic glutamate levels through a wide
range of mechanisms, and an initial trial found it to be eective
in treating negative symptoms in schizophrenia, potentially by
altering striatocortical connectivity254,255. Similarly, lamotrigine
inhibits glutamate release via inhibition of several ion channels,
and attenuates the psychotomimetic eects of ketamine256. La-
motrigine has also shown ecacy as an adjunctive medication
for clozapine-resistant schizophrenia, although studies to date
are small and ndings inconsistent257.
Neuroimaging studies have suggested that treatment-resist-
ant schizophrenia may not show the dopaminergic dysfunction
seen in treatment-responsive schizophrenia84,231,258, and that
glutamatergic abnormalities may be of greater pathophysiologi-
cal relevance in those cases of schizophrenia which do not re-
spond to antipsychotic medications259,260. Supporting this view,
there is evidence that cortical glutamate levels are higher in pa-
tients with treatment-resistant schizophrenia relative to respon-
sive patients261. One of the reasons for unsuccessful clinical trials
may therefore be that glutamate modulating treatments are only
of signicant benet in a subgroup of patients.
OUTSTANDING QUESTIONS AND FUTURE
DIRECTIONS
ere are a number of outstanding issues when it comes to
understanding the role of dopamine and glutamate in schizo-
phrenia. In the case of glutamate, it is not possible to separate
extra- and intracellular compartments using MRS, and we can-
not accurately probe receptors and synaptic glutamate levels in
vivo. As a result, it is not currently possible to precisely character-
ize the nature of glutamate dysfunction in schizophrenia. It re-
mains unclear whether synaptic glutamate levels are abnormal,
whether receptors are altered, and where any alterations might
be localized within the brain.
Because of this, it is not clear whether treatments should aim
to reduce synaptic glutamate levels or augment glutamatergic
neurotransmission150. is likely contributes to the fact that to
date no glutamate modulating agents exist that demonstrate un-
equivocal ecacy in schizophrenia.
ere is a need for radioligands with reliable binding at the
NMDA receptor to allow for investigation of receptor abnormali-
ties in schizophrenia. PET ligands for other proteins involved in
glutamatergic signalling, such as AMPA receptors, enzymes in-
volved in glutamate synthesis and metabolism, and the kynure-
nine pathway, would also represent a considerable advance. In
the meantime, other methods, such as functional MRS, 11C-MRS,
GluCEST and 7T 1H-MRS, may advance our understanding by
allowing more precise inferences regarding the nature of gluta-
matergic abnormalities in schizophrenia.
Imaging studies have provided more information when it
comes to the dopamine system. However, several questionsre-
main unanswered, such as the nature of cortical dopamine
function in schizophrenia, whether a cortical hypodopaminer-
gic state co-exists with the striatal hyperdopaminergic condi-
tion, and how dopaminergic dysfunction evolves across illness
course.
The improved resolution of PET cameras has allowed for
greater anatomical precision in identifying the locus of dopa-
minergic dysfunction in schizophrenia. Early hypotheses had
suggested that this dysfunction might be characterized by aber-
rant mesolimbic function. However, the use of new PET cameras
has demonstrated that the greatest patient-control dierences
are in the dorsal striatum83. e resolution of PET is still relatively
coarse, however, and this limits the precision with which infer-
ences can be made about which specic neuron groups are af-
fected.
In addition to advances in hardware, further progress may be
made by employing novel methods, such as super-resolution
techniques, in which multiple low-resolution images are com-
bined to create a high-resolution image, or deep learning meth-
ods, where anatomical information from an MRI scan is used to
help improve the resolution of the PET image262,263. e limited
resolution of in vivo imaging techniques means that it is dicult
to directly test circuit level hypotheses regarding the nature of
disrupted glutamate-dopamine interactions in schizophrenia.
Instead, hypotheses regarding circuit interactions are largely
based on preclinical, post-mortem and pharmacological studies,
but await direct testing in patients.
Translation of research ndings to clinically useful applica-
tions is not straightforward, and compounds acting though mech-
anisms other than dopamine receptor antagonism have not
consistently shown ecacy in clinical trials. However, there ex-
ists a range of novel mechanisms for manipulation of both glu-
tamate and dopamine signalling that show potential and await
clinical testing. As discussed above, it may be that certain treat-
ments are only of benet in specic subgroups of patients, and
clinical benet may therefore be optimized by stratifying partici-
pants on the basis of underlying neurobiology231,259. Given the
coarseness of current clinical measures264, the development of
imaging biomarkers to evaluate treatment eects at a neurobio-
logical level may assist in moving the eld forward265.
e non-linearity inherent to neural signalling within a com-
plex network means that, even if one disregards potential incon-
sistencies between studies, a coherent integration of existing
ndings is challenging. e combination of large datasets across
illness phases, biophysical networks models to link molecular
pathology to the macroscale dysfunction observed with neuro-
imaging, and carefully designed experiments to test and nesse
28 World Psychiatry 19:1 - Februar y 2020
these models is one route to integrating what at times appears to
be a disparate collection of ndings266,267.
CONCLUSIONS
e hypothesis that dopamine signalling is altered in schizo-
phrenia is supported by animal studies, post-mortem research,
and the clinical eects of drugs that either block or accentuate
dopaminergic neurotransmission. In addition, over the past 25
years, substantial evidence has accumulated from PET studies that
there is increased dopamine synthesis and release capacity in
schizophrenia, that is greatest within the dorsal striatum.
Genetic ndings do not provide strong support for the idea
that dopaminergic dysregulation is a primary abnormality. Rath-
er, it appears that the dopaminergic dysfunction is more likely
to develop downstream of abnormalities in other systems, in-
cluding the glutamatergic system. It also appears that environ-
mental factors may play a signicant role in the development of
dopaminergic dysregulation. Dopamine antagonists remain the
mainstay for pharmacological treatment of schizophrenia, but
there is increasing evidence that these are not eective for all pa-
tients.
Evidence for glutamate playing a role in the pathophysiology
of schizophrenia initially came from the psychotomimetic eects
of NMDA antagonists. While preclinical and post-mortem nd-
ings are consistent with this hypothesis, there is limited support
from imaging studies. However, in contrast to dopamine, recent
genetic ndings do provide support for the view that glutamater-
gic abnormalities may play a major role in schizophrenia patho-
physiology. However, progress is hampered by the challenges
involved in precisely characterizing the system in vivo, and, while
a wide range of glutamate modulating agents have been investi-
gated, none have clear clinical ecacy.
Despite the limitations described, as regards both treatment
ecacy and direct evidence for dysfunction, the dopamine and
glutamate hypotheses of schizophrenia remain inuential and
relevant. is is not least because, as recent data demonstrate,
they possess the exibility to accommodate new ndings, and to
provide ongoing potential avenues for the development of novel
treatments.
ACKNOWLEDGEMENTS
R.A. McCutcheon’s work is funded by a Wellcome trust grant (no. 200102/Z/
15/Z) and UK National Institute for Health Research (NIHR) fellowships.
REFERENCES
1. McCutcheon RA, Marques TR, Howes OD. Schizophrenia: an overview.
JAMA Psychiatry (in press).
2. Hjorthøj C, Stürup AE, McGrath JJ et al. Years of potential life lost and life
expectancy in schizophrenia: a systematic review and meta-analysis. Lancet
Psychiatry 2017;4:295-301.
3. van Rossum JM. e signicance of dopamine-receptor blockade for the
mechanism of action of neuroleptic drugs. Arch Int Pharmacodyn érapie
1966;160:492-4.
4. Kaar SJ, Natesan S, McCutcheon R et al. Antipsychotics: mechanisms un-
derlying clinical response and side-eects and novel treatment approaches
based on pathophysiology. Neuropharmacology (in press).
5. Carlsson A, Lindqvist M, Magnusson T. 3,4-Dihydroxyphenylalanine and
5-hydroxytryptophan as reserpine antagonists. Nature 1957;180:1200.
6. Featherstone RE, Kapur S, Fletcher PJ. e amphetamine-induced sensi-
tized state as a model of schizophrenia. Prog Neuropsychopharmacol Biol
Psychiatry 2007;31:1556-71.
7. Kellendonk C, Simpson EH, Polan HJ et al. Transient and selective overex-
pression of dopamine D2 receptors in the striatum causes persistent abnor-
malities in prefrontal cortex functioning. Neuron 2006;49:603-15.
8. Petty A, Cui X, Tesiram Y et al. Enhanced dopamine in prodromal schizo-
phrenia (EDiPS): a new animal model of relevance to schizophrenia. NPJ
Schizophr 2019;5:6.
9. Grace AA. Dopamine system dysregulation and the pathophysiology of
schizophrenia: insights from the methylazoxymethanol acetate model. Biol
Psychiatry 2017;81:5-8.
10. Lapiz MDS, Fulford A, Muchimapura S et al. Inuence of postweaning so-
cial isolation in the rat on brain development, conditioned behavior, and
neurotransmission. Neurosci Behav Physiol 2003;33:13-29.
11. Bowers MB. Central dopamine turnover in schizophrenic syndromes. Arch
Gen Psychiatry 1974;31:50-4.
12. Kahn RS, Harvey PD, Davidson M et al. Neuropsychological correlates of
central monoamine function in chronic schizophrenia: relationship between
CSF metabolites and cognitive function. Schizophr Res 1994;11:217-24.
13. Gattaz WF, Riederer P, Reynolds GP et al. Dopamine and noradrenalin in the
cerebrospinal uid of schizophrenic patients. Psychiatry Res 1983;8:243-50.
14. Gattaz WF, Gasser T, Beckmann H. Multidimensional analysis of the con-
centrations of 17 substances in the CSF of schizophrenics and controls. Biol
Psychiatry 1985;20:360-6.
15. Frecska E, Perényi A, Bagdy G et al. CSF dopamine turnover and positive
schizophrenic symptoms after withdrawal of long-term neuroleptic treat-
ment. Psychiatry Res 1985;16:221-6.
16. Bagdy G, Perényi A, Frecska E et al. Decrease in dopamine, its metabolites
and noradrenaline in cerebrospinal uid of schizophrenic patients after
withdrawal of long-term neuroleptic treatment. Psychopharmacology 1985;
85:62-4.
17. Maas JW, Bowden CL, Miller AL et al. Schizophrenia, psychosis, and ce-
rebral spinal uid homovanillic acid concentrations. Schizophr Bull 1997;
23:147-54.
18. van Kammen DP, Kelley M. Dopamine and norepinephrine activity in
schizophrenia. Schizophr Res 1991;4:173-91.
19. Post RM, Fink E, Carpenter WT et al. Cerebrospinal uid amine metabolites
in acute schizophrenia. Arch Gen Psychiatry 1975;32:1063-9.
20. Nybäck H, Berggren BM, Hindmarsh T et al. Cerebroventricular size and
cerebrospinal uid monoamine metabolites in schizophrenic patients and
healthy volunteers. Psychiatry Res 1983;9:301-8.
21. Robinson DG, Woerner MG, Alvir JM et al. Predictors of treatment response
from a rst episode of schizophrenia or schizoaective disorder. Am J Psy-
chiatry 1999;156:544-9.
22. Lee T, Seeman P. Elevation of brain neuroleptic/dopamine receptors in
schizophrenia. Am J Psychiatry 1980;137:191-7.
23. Zakzanis KK, Hansen KT. Dopamine D2 densities and the schizophrenic
brain. Schizophr Res 1998;32:201-6.
24. Silvestri S, Seeman MV, Negrete J-C et al. Increased dopamine D2 receptor
binding after long-term treatment with antipsychotics in humans: a clinical
PET study. Psychopharmacology 2000;152:174-80.
25. Samaha A-N, Seeman P, Stewart J et al. “Breakthrough” dopamine super-
sensitivity during ongoing antipsychotic treatment leads to treatment fail-
ure over time. J Neurosci 2007;27:2979-86.
26. Mueller HT, Haroutunian V, Davis KL et al. Expression of the ionotropic glu-
tamate receptor subunits and NMDA receptor-associated intracellular pro-
teins in the substantia nigra in schizophrenia. Mol Brain Res 2004;121:60-9.
27. Ichinose H, Ohye T, Fujita K et al. Quantication of mRNA of tyrosine hy-
droxylase and aromatic L-amino acid decarboxylase in the substantia nigra
in Parkinson’s disease and schizophrenia. J Neural Transm - Park Dis De-
ment Sect 1994;8:149-58.
28. Purves-Tyson TD, Owens SJ, Rothmond DA et al. Putative presynaptic do-
pamine dysregulation in schizophrenia is supported by molecular evidence
from post-mortem human midbrain. Transl Psychiatry 2017;7:e1003.
29. Williams MR, Galvin K, O’Sullivan B et al. Neuropathological changes in the
substantia nigra in schizophrenia but not depression. Eur Arch Psychiatry
Clin Neurosci 2014;264:285-96.
World Psychiatry 19:1 - February 2020 29
30. Gandal MJ, Zhang P, Hadjimichael E et al. Transcriptome-wide isoform-level
dysregulation in ASD, schizophrenia, and bipolar disorder. Science 2018;
362:6420.
31. Seeman P, Lee T. Antipsychotic drugs: direct correlation between clinical po-
tency and presynaptic action on dopamine neurons. Science 1975;188:1217-9.
32. Creese I, Burt D, Snyder S. Dopamine receptor binding predicts clinical and
pharmacological potencies of antischizophrenic drugs. Science 1976;192:481-
3.
33. Huhn M, Nikolakopoulou A, Schneider-oma J et al. Comparative ecacy
and tolerability of 32 oral antipsychotics for the acute treatment of adults
with multi-episode schizophrenia: a systematic review and network meta-
analysis. Lancet 2019;394:939-51.
34. Connell P. Amphetamine psychosis. BMJ 1957;5018:582.
35. Bell DS. e experimental reproduction of amphetamine psychosis. Arch
Gen Psychiatry 1973;29:35-40.
36. Moskovitz C, Moses H, Klawans HL. Levodopa-induced psychosis: a kin-
dling phenomenon. Am J Psychiatry 1978;135:669-75.
37. Voce A, Calabria B, Burns R et al. A systematic review of the symptom prole
and course of methamphetamine-associated psychosis: substance use and
misuse. Subst Use Misuse 2019;54:549-59.
38. Go DC, Coyle JT. e emerging role of glutamate in the pathophysiology
and treatment of schizophrenia. Am J Psychiatry 2001;158:1367-77.
39. Balu DT, Li Y, Puhl MD et al. Multiple risk pathways for schizophrenia con-
verge in serine racemase knockout mice, a mouse model of NMDA receptor
hypofunction. Proc Natl Acad Sci USA 2013;110:E2400-9.
40. Labrie V, Duy S, Wang W et al. Genetic inactivation of D-amino acid oxidase
enhances extinction and reversal learning in mice. Learn Mem 2009;16:28-37.
41. Carlson GC, Talbot K, Halene TB et al. Dysbindin-1 mutant mice implicate
reduced fast-phasic inhibition as a nal common disease mechanism in
schizophrenia. Proc Natl Acad Sci USA 2011;108:E962-70.
42. Kim JS, Kornhuber HH, Schmid-Burgk W et al. Low cerebrospinal uid glu-
tamate in schizophrenic patients and a new hypothesis on schizophrenia.
Neurosci Lett 1980;20:379-82.
43. Go DC, Wine L. Glutamate in schizophrenia: clinical and research impli-
cations. Schizophr Res 1997;27:157-68.
44. Plitman E, Iwata Y, Caravaggio F et al. Kynurenic acid in schizophrenia: a
systematic review and meta-analysis. Schizophr Bull 2017;43:764-77.
45. Hu W, Macdonald ML, Elswick DE et al. e glutamate hypothesis of schizo-
phrenia: evidence from human brain tissue studies. Ann NY Acad Sci 2015;
1338:38-57.
46. Javitt D, Zukin S. Recent advances in the phencyclidine model of schizo-
phrenia. Am J Psychiatry 1991;148:1301-8.
47. Stone JM, Pepper F, Fam J et al. Glutamate, N-acetyl aspartate and psychotic
symptoms in chronic ketamine users. Psychopharmacology 2014;231:2107-
16.
48. Fan N, Xu K, Ning Y et al. Proling the psychotic, depressive and anxiety
symptoms in chronic ketamine users. Psychiatry Res 2016;237:311-5.
49. Cassidy CM, Zucca FA, Girgis RR et al. Neuromelanin-sensitive MRI as a
noninvasive proxy measure of dopamine function in the human brain. Proc
Natl Acad Sci USA 2019;116:5108-17.
50. Urban NBL, Slifstein M, Meda S et al. Imaging human reward processing
with positron emission tomography and functional magnetic resonance im-
aging. Psychopharmacology 2012;221:67-77.
51. Radua J, Schmidt A, Borgwardt SJ et al. Ventral striatal activation during reward
processing in psychosis: a neurofunctional meta-analysis. JAMA Psychiatry
2016;72:1243-51.
52. Howes OD, Kambeitz J, Stahl D et al. e nature of dopamine dysfunction in
schizophrenia and what this means for treatment. Arch Gen Psychiatry 2012;
69:776-86.
53. Abi-Dargham A, van de Giessen E, Slifstein M et al. Baseline and ampheta-
mine-stimulated dopamine activity are related in drug-naïve schizophrenic
subjects. Biol Psychiatry 2009;65:1091-3.
54. Kegeles LS, Abi-Dargham A, Frankle WG et al. Increased synaptic dopa-
mine function in associative regions of the striatum in schizophrenia. Arch
Gen Psychiatry 2010;67:231-9.
55. Kubota M, Nagashima T, Takano H et al. Anity states of striatal dopamine
D2 receptors in antipsychotic-free patients with schizophrenia. Int J Neu-
ropsychopharmacol 2017;20:928-35.
56. Mizrahi R, Addington J, Rusjan PM et al. Increased stress-induced dopa-
mine release in psychosis. Biol Psychiatry 2012;71:561-7.
57. Frankle WG, Paris J, Himes M et al. Amphetamine-induced striatal dopa-
mine release measured with an agonist radiotracer in schizophrenia. Biol
Psychiatry 2018;83:707-14.
58. Slifstein M, Abi-Dargham A. Is it pre- or postsynaptic? Imaging striatal do-
pamine excess in schizophrenia. Biol Psychiatry 2018;83:635-7.
59. Guo N, Guo W, Kralikova M et al. Impact of D2 receptor internalization on
binding anity of neuroimaging radiotracers. Neuropsychopharmacology
2010;35:806-17.
60. Sun W, Ginovart N, Ko F et al. In vivo evidence for dopamine-mediated in-
ternalization of D2-receptors after amphetamine: dierential ndings with
[3H]raclopride versus [3H]spiperone. Mol Pharmacol 2003;63:456-62.
61. Brugger SP, Angelescu I, Abi-Dargham A et al. Heterogeneity of striatal do-
pamine function in schizophrenia: meta-analysis of variance. Biol Psychia-
try (in press).
62. Stenkrona P, Matheson GJ, Halldin C et al. D1-dopamine receptor availabil-
ity in rst-episode neuroleptic naive psychosis patients. Int J Neuropsycho-
pharmacol 2019;22:415-25.
63. Kambeitz J, Abi-Dargham A, Kapur S et al. Alterations in cortical and extra-
striatal subcortical dopamine function in schizophrenia: systematic review
and meta-analysis of imaging studies. Br J Psychiatry 2014;204:420-9.
64. Abi-Dargham A, Mawlawi O, Lombardo I et al. Prefrontal dopamine D1 re-
ceptors and working memory in schizophrenia. J Neurosci 2002;22:3708-
19.
65. Abi-Dargham A, Xu X, ompson JL et al. Increased prefrontal cortical D₁
receptors in drug naive patients with schizophrenia: a PET study with [11C]
NNC112. J Psychopharmacol 2012;26:794-805.
66. Kosaka J, Takahashi H, Ito H et al. Decreased binding of [11C]NNC112 and
[11C]SCH23390 in patients with chronic schizophrenia. Life Sci 2010;86:814-
8.
67. Hirvonen J, Van Erp TGM, Huttunen J et al. Brain dopamine D1 receptors in
twins discordant for schizophrenia. Am J Psychiatry 2006;163:1747-53.
68. Okubo Y, Suhara T, Suzuki K et al. Decreased prefrontal dopamine D1 re-
ceptors in schizophrenia revealed by PET. Nature 1997;385:634-6.
69. Karlsson P, Farde L, Halldin C et al. PET study of D(1) dopamine receptor
bind ing in neuroleptic-naive patients with schizophrenia. Am J Psychiatry
2002;159:761-7.
70. Poels EMP, Girgis RR, ompson JL et al. In vivo binding of the dopamine-1
receptor PET tracers [11C]NNC112 and [11C]SCH23390: a comparison study
in individuals with schizophrenia. Psychopharmacology 2013;228:167-74.
71. Guo N, Hwang DR, Lo ES et al. Dopamine depletion and in vivo binding of
PET D1 receptor radioligands: implications for imaging studies in schizo-
phrenia. Neuropsychopharmacology 2003;28:1703-11.
72. Ekelund J, Slifstein M, Narendran R et al. In vivo DA D(1) receptor selectivity
of NNC 112 and SCH 23390. Mol Imaging Biol 2007;9:117-25.
73. Lidow MS, Elsworth JD, Goldman-Rakic PS. Down-regulation of the D1 and
D5 dopamine receptors in the primate prefrontal cortex by chronic treat-
ment with antipsychotic drugs. J Pharmacol Exp er 1997;281:597-603.
74. Artiges E, Leroy C, Dubol M et al. Striatal and extrastriatal dopamine trans-
porter availability in schizophrenia and its clinical correlates: a voxel-based
and high-resolution PET study. Schizophr Bull 2017;43:1134-42.
75. Arakawa R, Ichimiya T, Ito H et al. Increase in thalamic binding of [11C]PE2I
in patients with schizophrenia: a positron emission tomography study of
dopamine transporter. J Psychiatr Res 2009;43:1219-23.
76. Taylor SF, Koeppe RA, Tandon R et al. In vivo measurement of the vesicu-
lar monoamine transporter in schizophrenia. Neuropsychopharmacology
2000;23:667-75.
77. Zubieta JK, Taylor SF, Huguelet P et al. Vesicular monoamine transporter
concentrations in bipolar disorder type I, schizophrenia, and healthy sub-
jects. Biol Psychiatry 2001;49:110-6.
78. Zucker M, Valevski A, Weizman A et al. Increased platelet vesicular mono-
amine transporter density in adult schizophrenia patients. Eur Neuropsy-
chopharmacol 2002;12:343-7.
79. Egerton A, Mehta MA, Montgomery AJ et al. e dopaminergic basis of hu-
man behaviors: a review of molecular imaging studies. Neurosci Biobehav
Rev 2009;33:1109-32.
80. Dewey SL, Logan J, Wolf AP et al. Amphetamine induced decreases in
(18F)-N-methylspiroperidol binding in the baboon brain using positron
emission tomography (PET). Synapse 1991;7:324-7.
81. Kumakura Y, Cumming P. PET studies of cerebral levodopa metabolism: a
review of clinical ndings and modeling approaches. Neuroscientist 2009;
15:635-50.
82. Sawilowsky SS. New eect size rules of thumb. J Mod Appl Stat Methods
2009;8:597-9.
83. McCutcheon R, Beck K, Jauhar S et al. Dening the locus of dopaminergic
dysfunction in schizophrenia: a meta-analysis and test of the mesolimbic
hypothesis. Schizophr Bull 2018;44:1301-11.
30 World Psychiatry 19:1 - Februar y 2020
84. Demjaha A, Murray RM, McGuire PK et al. Dopamine synthesis capacity
in patients with treatment-resistant schizophrenia. Am J Psychiatry 2012;
169:1203-10.
85. Abi-Dargham A, Gil R, Krystal J et al. Increased striatal dopamine transmis-
sion in schizophrenia: conrmation in a second cohort. Am J Psychiatry 1998;
155:761-7.
86. Jauhar S, Nour MM, Veronese M et al. A test of the transdiagnostic dopamine
hypothesis of psychosis using positron emission tomographic imaging in
bipolar aective disorder and schizophrenia. JAMA Psychiatry 2017;74:1206-13.
87. Hietala J, Syvälahti E, Vilkman H et al. Depressive symptoms and presynap-
tic dopamine function in neuroleptic-naive schizophrenia. Schizophr Res
1999;35:41-50.
88. Avram M, Brandl F, Cabello J et al. Reduced striatal dopamine synthesis ca-
pacity in patients with schizophrenia during remission of positive symptoms.
Brain 2019;142:1813-26.
89. Egerton A, Demjaha A, McGuire P et al. e test-retest reliability of 18F-
DOPA PET in assessing striatal and extrastriatal presynaptic dopaminergic
function. Neuroimage 2010;50:524-31.
90. Howes OD, Williams M, Ibrahim K et al. Midbrain dopamine function in
schizophrenia and depression: a post-mortem and positron emission to-
mographic imaging study. Brain 2013;136:3242-51.
91. Kumakura Y, Cumming P, Vernaleken I et al. Elevated [18F]uorodopamine
turnover in brain of patients with schizophrenia: an [18F]uorodopa/posi-
tron emission tomography study. J Neurosci 2007;27:8080-7.
92. Elkashef A, Doudet D, Bryant T. 6-(18)F-DOPA PET study in patients with
schizophrenia. Psychiatry Res Neuroimaging 2000;100:1-11.
93. Hall H, Sedvall G, Magnusson O et al. Distribution of D1- and D2-dopamine
receptors, and dopamine and its metabolites in the human brain. Neuropsy-
chopharmacology 1994;11:245-56.
94. Hernaus D, Mehta MA. Prefrontal cortex dopamine release measured in vivo
with positron emission tomography: implications for the stimulant paradigm.
Neuroimage 2016;142:663-7.
95. Slifstein M, Van De Giessen E, Van Snellenberg J et al. Decits in prefrontal
cortical and extrastriatal dopamine release in schizophrenia a positron emis-
sion tomographic functional magnetic resonance imaging study. JAMA Psy-
chiatry 2015;72:316-24.
96. Schifani C, Tseng H-H, Kenk M et al. Cortical stress regulation is disrupted in
schizophrenia but not in clinical high risk for psychosis. Brain 2018;141:2213-
24.
97. Rao N, Northo G, Tagore A et al. Impaired prefrontal cortical dopamine re-
lease in schizophrenia during a cognitive task : a [11C]FLB 457 positron emis-
sion tomography study. Schizophr Bull 2019;45:670-9.
98. Hernaus D, Collip D, Kasanova Z et al. No evidence for attenuated stress-in-
duced extrastriatal dopamine signaling in psychotic disorder. Transl Psychia-
try 2015;5:e547.
99. Tseng H-H, Watts JJ, Kiang M et al. Nigral stress-induced dopamine release
in clinical high risk and antipsychotic-naïve schizophrenia. Schizophr Bull
2018;44:542-51.
100. Abi-Dargham A, Kegeles LS, Zea-Ponce Y et al. Striatal amphetamine-in-
duced dopamine release in patients with schizotypal personality disorder
studied with single photon emission computed tomography and [123I]io-
dobenzamide. Biol Psychiatry 2004;55:1001-06.
101. Howes OD, Montgomery AJ, Asselin M-C et al. Elevated striatal dopamine
function linked to prodromal signs of schizophrenia. Arch Gen Psychiatry
2009;66:13-20.
102. Egerton A, Chaddock CA, Winton-Brown TT et al. Presynaptic striatal do-
pamine dysfunction in people at ultra-high risk for psychosis: ndings in a
second cohort. Biol Psychiatry 2013;74:106-12.
103. Howes OD, Bonoldi I, McCutcheon RA et al. Glutamatergic and dopaminer-
gic function and the relationship to outcome in people at clinical high risk
of psychosis: a multi-modal PET-magnetic resonance brain imaging study.
Neuropsychopharmacology (in press).
104. Howes O, Bose S, Turkheimer FE et al. Dopamine synthesis capacity before
onset of psychosis: a prospective [18F]-DOPA PET imaging study. Am J Psy-
chiatry 2011;168:1311-7.
105. Howes OD, Shotbolt P, Bloomeld M et al. Dopaminergic function in the
psychosis spectrum: an [18F]-DOPA imaging study in healthy individuals
with auditory hallucinations. Schizophr Bull 2013;39:807-14.
106. Shotbolt P, Stokes PR, Owens SF et al. Striatal dopamine synthesis capacity
in twins discordant for schizophrenia. Psychol Med 2011;41:2331-8.
107. van Duin EDA , Kasanova Z, Hernaus D et al. Striatal dopamine release and
impaired reinforcement learning in adults with 22q11.2 deletion syndrome.
Eur Neuropsychopharmacol 2018;28:732-42.
108. Huttunen J, Heinimaa M, Svirskis T et al. Striatal dopamine synthesis in rst-
degree relatives of patients with schizophrenia. Biol Psychiatry 2008;63:114-
7.
109. Reith J, Benkelfat C, Sherwin A et al. Elevated dopa decarboxylase activity in
living brain of patients with psychosis. Proc Natl Acad Sci USA 1994;91:11651-
4.
110. Vingerhoets C, Bloemen OJN, Boot E et al. Dopamine in high-risk popula-
tions: a comparison of subjects with 22q11.2 deletion syndrome and subjects
at ultra high-risk for psychosis. Psychiatry Res Neuroimaging 2018;272:65-70.
111. Brunelin J, d’Amato T, van Os J et al. Increased left striatal dopamine trans-
mission in unaected siblings of schizophrenia patients in response to a-
cute metabolic stress. Psychiatry Res 2010;181:130-5.
112. Lee KJ, Lee JS, Kim SJ et al. Loss of asymmetry in D2 receptors of putamen in
unaected family members at increased genetic risk for schizophrenia. Acta
Psychiatr Scand 2008;118:200-8.
113. Hirvonen J, van Erp TGM, Huttunen J et al. Striatal dopamine D1 and D2
receptor balance in twins at increased genetic risk for schizophrenia. Psy-
chiatry Res 2006;146:13-20.
114. Fouriezos G, Hansson P, Wise RA . Neuroleptic-induced attenuation of brain
stimulation reward in rats. J Comp Physiol Psychol 1978;92:661-71.
115. Schultz W, Dayan P, Montague PR. A neural substrate of prediction and re-
ward. Science 1997;275:1593-9.
116. Lerner TN, Shilyansky C, Davidson TJ et al. Intact-brain analyses reveal dis-
tinct information carried by SNc dopamine subcircuits. Cell 2015;162:635-
47.
117. Menegas W, Babayan BM, Uchida N et al. Opposite initialization to novel
cues in dopamine signaling in ventral and posterior striatum in mice. Elife
2017;6:e21886.
118. Menegas W, Akiti K, Amo R et al. Dopamine neurons projecting to the pos-
terior striatum reinforce avoidance of threatening stimuli. Nat Neurosci
2018;21:1421-30.
119. Groessl F, Munsch T, Meis S et al. Dorsal tegmental dopamine neurons gate
associative learning of fear. Nat Neurosci 2018;21:952-62.
120. Matsumoto M, Hikosaka O. Two types of dopamine neuron distinctly con-
vey positive and negative motivational signals. Nature 2009;459:837-41.
121. Miller R. Schizophrenic psychology, associative learning and the role of
forebrain dopamine. Med Hypotheses 1976;2:203-11.
122. Kapur S. Psychosis as a state of aberrant salience: a framework linking biol-
ogy, phenomenology, and pharmacology in schizophrenia. Am J Psychiatry
2003;160:13-23.
123. McCutcheon RA, Abi-Dargham A, Howes OD. Schizophrenia, dopamine
and the striatum: from biology to symptoms. Trends Neurosci 2019;42:205-
20.
124. Sterzer P, Adams RA, Fletcher P et al. e predictive coding account of psy-
chosis. Biol Psychiatry 2018;84:634-43.
125. Corlett PR, Kr ystal JH, Taylor JR et al. Why do delusions persist? Front Hum
Neurosci 2009;3:12.
126. Powers AR, Mathys C, Corlett PR. Pavlovian conditioning-induced halluci-
nations result from overweighting of perceptual priors. Science 2017;357:596-
600.
127. Cassidy CM, Balsam PD, Weinstein JJ et al. A perceptual inference mecha-
nism for hallucinations linked to striatal dopamine. Curr Biol 2018;28:503-
14.
128. Nour MM, Dahoun T, Schwartenbeck P et al. Dopaminergic basis for signal-
ing belief updates, but not surprise, and the link to paranoia. Proc Natl Acad
Sci USA 2018;115:E10167-76
129. McCutcheon RA, Bloomeld MAP, Dahoun T et al. Chronic psychosocial
stressors are associated with alterations in salience processing and cortico-
striatal connectivity. Schizophr Res 2019;213:56-64.
130. Uddin LQ. Salience processing and insular cortical function and dysfunc-
tion. Nat Rev Neurosci 2015;16:55-61.
131. Palaniyappan L, Simmonite M, White TP et al. Neural primacy of the sali-
ence processing system in schizophrenia. Neuron 2013;79:814-28.
132. Menon V, Uddin LQ. Saliency, switching, attention and control: a network
model of insula function. Brain Struct Funct 2010;214:655-67.
133. Roman JL, Tanner AS, Eryilmaz H et al. Dopamine D1 signaling organizes
network dynamics underlying working memory. Sci Adv 2016;2:e1501672.
134. Braun U, Harneit A, Pergola G et al. Brain state stability during working mem-
ory is explained by network control theory, modulated by dopamine D1/D2
receptor function, and diminished in schizophrenia. bioRxiv 2019:679670.
135. McCutcheon R, Nour MM, Dahoun T et al. Mesolimbic dopamine function
is related to salience network connectivity: an integrative positron emission
tomography and magnetic resonance study. Biol Psychiatry 2019;85:368-78.
World Psychiatry 19:1 - February 2020 31
136. Maia TV, Frank MJ. From reinforcement learning models to psychiatric and
neurological disorders. Nat Neurosci 2011;14:154-62.
137. Krystal JH, Anticevic A, Yang GJ et al. Impaired tuning of neural ensembles
and the pathophysiology of schizophrenia: a translational and computa-
tional neuroscience perspective. Biol Psychiatry 2017;81:874-85.
138. Schultz W. Predictive reward signal of dopamine neurons. J Neurophysiol
1998;80:1-27.
139. Maia TV, Frank MJ. An integrative perspective on the role of dopamine in
schizophrenia. Biol Psychiatry 2017;81:52-66.
140. Strauss GP, Frank MJ, Waltz JA et al. Decits in positive reinforcement learn-
ing and uncertainty-driven exploration are associated with distinct aspects
of negative symptoms in schizophrenia. Biol Psychiatry 2011;69:424-31.
141. Gold JM, Waltz JA, Matveeva TM et al. Negative symptoms in schizophrenia
result from a failure to represent the expected value of rewards: behavioral
and computational modeling evidence. Arch Gen Psychiatry 2012;69:129-
38.
142. Krystal JH, D’Souza DC, Gallinat J et al. e vulnerability to alcohol and
substance abuse in individuals diagnosed with schizophrenia. Neurotox
Res 2006;10:235-52.
143. MacCabe JH, Wicks S, Löfving S et al. Decline in cognitive performance be-
tween ages 13 and 18 years and the risk for psychosis in adulthood: a Swed-
ishlongitudinal cohort study in males. JAMA Psychiatry 2013;70:261-70.
144. Keefe RSE, Fenton WS. How should DSM-V criteria for schizophrenia in-
clude cognitive impairment? Schizophr Bull 2007;33:912-20.
145. Green MF, Horan PW, Lee J. Nonsocial and social cognition in schizophrenia:
current evidence and future directions. World Psychiatry 2019;18:146-61.
146. Kelly S, Guimond S, Lyall A et al. Neural correlates of cognitive decits across
developmental phases of schizophrenia. Neurobiol Dis 2019;131:104353.
147. Arnsten AFT. Catecholamine modulation of prefrontal cortical cognitive
function. Trends Cogn Sci 1998;2:436-47.
148. Schobel SA, Chaudhury NH, Khan UA et al. Imaging patients with psychosis
and a mouse model establishes a spreading pattern of hippocampal dys-
function and implicates glutamate as a driver. Neuron 2013;78:81-93.
149. Lieberman JA, Girgis RR, Brucato G et al. Hippocampal dysfunction in the
pathophysiology of schizophrenia: a selective review and hypothesis for
early detection and intervention. Mol Psychiatry 2018;23:1764-72.
150. Moghaddam B, Javitt D. From revolution to evolution: the glutamate hy-
pothesis of schizophrenia and its implication for treatment. Neuropsycho-
pharmacology 2012;37:4-15.
151. Malhotra AK, Pinals DA, Adler CM et al. Ketamine-induced exacerbation of
psychotic symptoms and cognitive impairment in neuroleptic-free schizo-
phrenics. Neuropsychopharmacology 1997;17:141-50.
152. Al-Diwani A, Handel A, Townsend L et al. e psychopathology of NMDAR-
antibody encephalitis in adults: a systematic review and phenotypic analy-
sis of individual patient data. Lancet Psychiatry 2019;6:235-46.
153. Krystal JH, Karper LP, Seibyl JP et al. Subanesthetic eects of the noncom-
petitive NMDA antagonist, ketamine, in humans. Psychotomimetic, per-
ceptual, cognitive, and neuroendocrine responses. Arch Gen Psychiatry
1994;51:199-214.
154. Carhart-Harris RL, Brugger S, Nutt DJ et al. Psychiatry’s next top model:
cause for a re-think on drug models of psychosis and other psychiatric dis-
orders. J Psychopharmacol 2013;27:771-8.
155. Cheng WJ, Chen CH, Chen CK et al. Similar psychotic and cognitive prole
between ketamine dependence with persistent psychosis and schizophre-
nia. Schizophr Res 2018;199:313-8.
156. Poels EMP, Kegeles LS, Kantrowitz JT et al. Imaging glutamate in schizo-
phrenia: review of ndings and implications for drug discovery. Mol Psy-
chiatry 2014;19:20-9.
157. Bak LK, Schousboe A, Waagepetersen HS. e glutamate/GABA-glutamine
cycle: aspects of transport, neurotransmitter homeostasis and ammonia
transfer. J Neurochem 2006;98:641-53.
158. McKenna MC. e glutamate-glutamine cycle is not stoichiometric: fates of
glutamate in brain. J Neurosci Res 2007;85:3347-58.
159. Merritt K, Egerton A, Kempton MJ et al. Nature of glutamate alterations in
schizophrenia. JAMA Psychiatry 2016;52:998-1007.
160. Reid MA, Salibi N, White DM et al. 7T proton magnetic resonance spectros-
copy of the anterior cingulate cortex in rst-episode schizophrenia. Schizo-
phr Bull 2019;45:180-9.
161. Wang AM, Pradhan S, Coughlin JM et al. Assessing brain metabolism with
7-T proton magnetic resonance spectroscopy in patients with rst-episode
psychosis. JAMA Psychiatry 2019;76:314-23.
162. Kumar J, Liddle EB, Fernandes CC et al. Glutathione and glutamate in schiz-
ophrenia: a 7T MRS study. Mol Psychiatry (in press).
163. Marsman A, Mandl RCW, Klomp DWJ et al. GABA and glutamate in schizo-
phrenia: a 7T 1H-MRS study. Neuroimage Clin 2014;6:398-407.
164. Posporelis S, Coughlin JM, Marsman A et al. Decoupling of brain tempera-
ture and glutamate in recent onset of schizophrenia: a 7T proton magnetic
resonance spectroscopy study. Biol Psychiatry Cogn Neurosci Neuroimag-
ing 2018;3:248-54.
165. Brandt AS, Unschuld PG, Pradhan S et al. Age-related changes in anterior cin-
gulate cortex glutamate in schizophrenia: a 1H MRS study at 7 Tesla. Schizo-
phr Res 2016;172:101-5.
166. Taylor R, Neufeld RWJ, Schaefer B et al. Functional magnetic resonance
spectroscopy of glutamate in schizophrenia and major depressive disorder:
anterior cingulate activity during a color-word Stroop task. NPJ Schizophr
2015;1:15028.
167. Jelen LA, King S, Mullins PG et al. Beyond static measures: a review of func-
tional magnetic resonance spectroscopy and its potential to investigate dy-
namic glutamatergic abnormalities in schizophrenia. J Psychopharmacol
2018;32:497-508.
168. Chiappelli J, Shi Q, Wijtenburg SA et al. Glutamatergic response to heat pain
stress in schizophrenia. Schizophr Bull 2018;44:886-95.
169. Akkus F, Treyer V, Ametamey SM et al. Metabotropic glutamate receptor 5
neuroimaging in schizophrenia. Schizophr Res 2017;183:95-101.
170. Fu H, Chen Z, Josephson L et al. Positron emission tomography (PET) ligand
development for ionotropic glutamate receptors: challenges and opportu-
nities for radiotracer targeting N-methyl-d-aspartate (NMDA), α-amino-3-
hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA), and kainate receptors.
J Med Chem 2019;62:403-19.
171. Mason G, Petersen K, de Graaf RA et al. Measurements of the anaplerotic
rate in the human cerebral cortex using 13C magnetic resonance spectros-
copy and [1-13 C] and [2-13C] glucose. J Neurochem 2007;100:73-86.
172. Abdallah CG, De Feyter HM, Averill LA et al. e eects of ketamine on
prefrontal glutamate neurotransmission in healthy and depressed subjects.
Neuropsychopharmacology 2018;43:2154-60.
173. Sibson NR, Dhankhar A, Mason GF et al. Stoichiometric coupling of brain
glucose metabolism and glutamatergic neuronal activity. Proc Natl Acad Sci
USA 1998;95:316-21.
174. Cai K, Haris M, Singh A et al. Magnetic resonance imaging of glutamate. Nat
Med 2012;18:302-6.
175. Roalf DR, Nanga RPR, Rupert PE et al. Glutamate imaging (GluCEST) re-
veals lower brain GluCEST contrast in patients on the psychosis spectrum.
Mol Psychiatry 2017;22:1298-305.
176. Modinos G, S¸ ims¸ek F, Azis M et al. Prefrontal GABA levels, hippocampal
resting perfusion and the risk of psychosis. Neuropsychopharmacology
2018;43:262-59.
177. Liemburg E, Sibeijn-Kuiper A, Bais L et al. Prefrontal NAA and Glx lev-
els in dierent stages of psychotic disorders: a 3T 1H-MRS study. Sci Rep
2016;6:1-8.
178. Da Silva T, Hazi S, Rusjan PM et al. GABA levels and TSPO expression in
people at clinical high risk for psychosis and healthy volunteers: a PET-MRS
study. J Psychiatry Neurosci 2019;44:111-9.
179. Hädel S, Wirth C, Rapp M et al. Eects of age and sex on the concentrations
of glutamate and glutamine in the human brain. J Magn Reson Imaging
2013;38:1480-7.
180. Tayoshi S, Sumitani S, Taniguchi K et al. Metabolite changes and gender dif-
ferences in schizophrenia using 3-Tesla proton magnetic resonance spec-
troscopy (1H-MRS). Schizophr Res 2009;108:69-77.
181. Grachev I, Apkarian A. Aging alters regional multichemical prole of the
human brain: an in vivo 1H-MRS study of young versus middle-aged sub-
jects. J Neurochem 2001;76:582-93.
182. Merritt K , McGuire P, Egerton A. Relationship between glutamate dysfunc-
tion and symptoms and cognitive function in psychosis. Front Psychiatry 2013;
4:151.
183. Lutkenho ES, van Erp TG, omas MA et al. Proton MRS in twin pairs dis-
cordant for schizophrenia. Mol Psychiatry 2010;15:308-18.
184. Tebartz Van Elst L, Valerius G, Büchert M et al. Increased prefrontal and
hippocampal glutamate concentration in schizophrenia: evidence from a
magnetic resonance spectroscopy study. Biol Psychiatry 2005;58:724-30.
185. Wang M, Yang Y, Wang CJ et al. NMDA receptors subserve persistent neu-
ronal ring during working memory in dorsolateral prefrontal cortex. Neu-
ron 2013;77:736-49.
186. Harris KD, Mrsic-Flogel TD. Cortical connectivity and sensory coding. Na-
ture 2013;503:51-8.
187. Buzsáki G, Watson BO. Brain rhythms and neural syntax: implications for ef-
cient coding of cognitive content and neuropsychiatric disease. Dialogues
32 World Psychiatry 19:1 - Februar y 2020
Clin Neurosci 2012;14:345-67.
188. Lin AC, Bygrave AM, De Calignon A et al. Sparse, decorrelated odor coding
in the mushroom body enhances learned odor discrimination. Nat Neuro-
sci 2014;17:559-68.
189. Murray JD, Anticevic A, Gancsos M et al. Linking microcircuit dysfunction
to cognitive impairment: eects of disinhibition associated with schizo-
phrenia in a cortical working memory model. Cereb Cortex 2014;24:859-72.
190. Uhlhaas PJ, Singer W. Abnormal neural oscillations and synchrony in schiz-
ophrenia. Nat Rev Neurosci 2010;11:100-13.
191. Lewis DA , Curley AA, Glausier JR et al. Cortical parvalbumin interneurons
and cognitive dysfunction in schizophrenia. Trends Neurosci 2012;35:57-
67.
192. Coyle J. Glutamate and schizophrenia: beyond the dopamine hypothesis.
Cell Mol Neurobiol 2006;26:365-84.
193. Pers TH, Timshel P, Ripke S et al. Comprehensive analysis of schizophrenia-
associated loci highlights ion channel pathways and biologically plausible
candidate causal genes. Hum Mol Genet 2015;25:1247-54.
194. Ma C, Gu C, Huo Y et al. e integrated landscape of causal genes and path-
ways in schizophrenia. Transl Psychiatry 2018;8:67.
195. Schizophrenia Working Group of the Psychiatric Genomics Consortium. Bio -
logical insights from 108 schizophrenia-associated genetic loci. Nature
2014;511:421-7.
196. Edwards AC, Bacanu S, Bigdeli TB et al. Evaluating the dopamine hypothesis
of schizophrenia in a large-scale genome-wide association study. Schizophr
Res 2016;176:136-40.
197. D’Ambrosio E, Dahoun T, Pardiñas AF et al. e eect of a genetic variant
at the schizophrenia associated AS3MT/BORCS7 locus on striatal dopamine
function: a PET imaging study. Psychiatry Res Neuroimaging 2019;291:34-41.
198. Dahoun T, Pardiñas AF, Veronese M et al. e eect of the DISC1 Ser704Cys
polymorphism on striatal dopamine synthesis capacity: an [18F]-DOPA
PET study. Hum Mol Genet 2018;27:3498-506.
199. Pocklington AJ, Rees E, Walters JTR et al. Novel ndings from CNVs impli-
cate inhibitory and excitatory signaling complexes in schizophrenia. Neu-
ron 2015;86:1203-14.
200. Timms AE, Dorschner MO, Wechsler J et al. Support for the N-methyl-D-
aspartate receptor hypofunction hypothesis of schizophrenia from exome
sequencing in multiplex families. JAMA Psychiatry 2013;70:582-90.
201. Brennand K, Savas JN, Kim Y et al. Phenotypic dierences in hiPSC NPCs
derived from patients with schizophrenia. Mol Psychiatry 2015;20:361-8.
202. Robicsek O, Karry R, Petit I et al. Abnormal neuronal dierentiation and mi-
tochondrial dysfunction in hair follicle-derived induced pluripotent stem
cells of schizophrenia patients. Mol Psychiatry 2013;18:1067-76.
203. Hook V, Brennand KJ, Kim Y et al. Human iPSC neurons display activity-
dependent neurotransmitter secretion: aberrant catecholamine levels in
schizophrenia neurons. Stem Cell Reports 2014;3:531-8.
204. Hartley BJ, Tran N, Ladran I et al. Dopaminergic dierentiation of schizo-
phrenia hiPSCs. Mol Psychiatry 2015;20:549-50.
205. Brennand KJ, Simone A, Jou J et al. Modelling schizophrenia using human
induced pluripotent stem cells. Nature 2011;473:221-5.
206. Wen Z, Nguyen HN, Guo Z et al. Synaptic dysregulation in a human iPS cell
model of mental disorders. Nature 2014;515:414-8.
207. Yu DX, Di Giorgio FP, Yao J et al. Modeling hippocampal neurogenesis using
human pluripotent stem cells. Stem Cell Reports 2014;2:295-310.
208. Pruessner JC, Champagne F, Meaney MJ et al. Dopamine release in response
to a psychological stress in humans and its relationship to early life maternal
care: a positron emission tomography study using [11C]raclopride. J Neurosci
2004;24:2825-31.
209. Vaessen T, Hernaus D, Myin-Germeys I et al. e dopaminergic response to
acute stress in health and psychopathology: a systematic review. Neurosci
Biobehav Rev 2015;56:241-51.
210. Egerton A, Howes OD, Houle S et al. Elevated striatal dopamine function in
immigrants and their children: a risk mechanism for psychosis. Schizophr
Bull 2017;43:293-301.
211. Egerton A, Valmaggia LR, Howes OD et al. Adversity in childhood linked to
elevated striatal dopamine function in adulthood. Schizophr Res 2016;176:
171-6.
212. Radua J, Ramella-Cravaro V, Ioannidis JPA et al. What causes psychosis? An
umbrella review of risk and protective factors. World Psychiatry 2018;17:49-
66.
213. Morgan C, Knowles G, Hutchinson G. Migration, ethnicity and psychoses:
evidence, models and future directions. World Psychiatry 2019;18:247-58.
214. Bloomeld MA, Morgan CJ, Egerton A et al. Dopaminergic function in can-
nabis users and its relationship to cannabis-induced psychotic symptoms.
Biol Psychiatry 2014;75:470-8.
215. Mizrahi R, Suridjan I, Kenk M et al. Dopamine response to psychosocial
stress in chronic cannabis users: a PET study with [11C]-+-PHNO. Neu-
ropsychopharmacology 2013;38:673-82.
216. Popoli M, Yan Z, McEwen BS et al. e stressed synapse: the impact of
stressand glucocorticoids on glutamate transmission. Nat Rev Neurosci 2012;
13:22-37.
217. Houtepen LC, Schür RR, Wijnen JP et al. Acute stress eects on GABA and
glutamate levels in the prefrontal cortex: a 7T 1H-magnetic resonance spec-
troscopy study. Neuroimage Clin 2017;14:195-200.
218. Colizzi M, McGuire P, Pertwee RG et al. Eect of cannabis on glutamate
signalling in the brain: a systematic review of human and animal evidence.
Neurosci Biobehav Rev 2016;64:359-81.
219. Guennewig B, Bitar M, Obiorah I et al. THC exposure of human iPSC neu-
rons impacts genes associated with neuropsychiatric disorders. Transl Psy-
chiatry 2018;8:89.
220. White TL, Monnig MA, Walsh EG et al. Psychostimulant drug eects on glu-
tamate, Glx, and creatine in the anterior cingulate cortex and subjective re-
sponse in healthy humans. Neuropsychopharmacology 2018;43:1498-509.
221. Egerton A, Bhachu A, Merritt K et al. Eects of antipsychotic administration
on brain glutamate in schizophrenia: a systematic review of longitudinal (1)
H-MRS studies. Front Psychiatry 2017;8:66.
222. Kokkinou M, Ashok A, Howes O. e eects of ketamine on dopaminergic
function: meta-analysis and review of the implications for neuropsychiatric
disorders. Mol Psychiatry 2018;23:59-69.
223. Krystal JH, D’Souza DC, Karper LP et al. Interactive eects of subanes-
thetic ketamine and haloperidol in healthy humans. Psychopharmacology
1999;145:193-204.
224. Gleich T, Deserno L, Lorenz RC et al. Prefrontal and striatal glutamate dif-
ferently relate to striatal dopamine: potential regulatory mechanisms of stri-
atal presynaptic dopamine function? J Neurosci 2015;35:9615-21.
225. Jauhar S, McCutcheon R, Borgan F et al. e relationship between cortical
glutamate and striatal dopamine in rst-episode psychosis: a cross-section-
al multimodal PET and magnetic resonance spectroscopy imaging study.
Lancet Psychiatry 2018;5:816-23.
226. Stone JM, Howes OD, Egerton A et al. Altered relationship between hip-
pocampal glutamate levels and striatal dopamine function in subjects at
ultra high risk of psychosis. Biol Psychiatry 2010;68:599-602.
227. Schwartz TL, Sachdeva S, Stahl SM. Glutamate neurocircuitry: theoretical
underpinnings in schizophrenia. Front Pharmacol 2012;3:195.
228. Chun S, Westmoreland JJ, Bayazitov IT et al. Specific disruption of tha-
lamic inputs to the auditory cortex in schizophrenia models. Science 2014;
344:1178-82.
229. Gao WJ, Krimer LS, Goldman-Rakic PS. Presynaptic regulation of recurrent
excitation by D1 receptors in prefrontal circuits. Proc Natl Acad Sci USA
2001;98:295-300.
230. Howes OD, McCutcheon R, Agid O et al. Treatment-resistant schizophre-
nia: Treatment Response and Resistance in Psychosis (TRRIP) Working
Group consensus guidelines on diagnosis and terminology. Am J Psychiatry
2017;174:216-29.
231. Jauhar S, Veronese M, Nour MM et al. Determinants of treatment re-
sponse in rst-episode psychosis: an 18F-DOPA PET study. Mol Psychiatry
2019;24:1502-12.
232. Tamminga CA, Schaer MH, Smith RC et al. Schizophrenic symptoms im-
prove with apomorphine. Science 1978;200:567-8.
233. Jauhar S, Veronese M, Rogdaki M et al. Regulation of dopaminergic func-
tion: an [18F]-DOPA PET apomorphine challenge study in humans. Transl
Psychiatry 2017;7:e1027.
234. Revel FG, Moreau J-L, Gainetdinov RR et al. TAAR1 activation modulates
monoaminergic neurotransmission, preventing hyperdopaminergic and
hypoglutamatergic activity. Proc Natl Acad Sci USA 2011;108:8485-90.
235. Foster DJ, Wilson JM, Remke DH et al. Antipsychotic-like eects of M4 posi-
tive allosteric modulators are mediated by CB2 receptor-dependent inhibi-
tion of dopamine release. Neuron 2016;91:1244-52.
236. Shekhar A, Potter WZ, Lightfoot J et al. Selective muscarinic receptor ago-
nist xanomeline as a novel treatment approach for schizophrenia. Am J Psy-
chiatry 2008;165:1033-9.
237. Heckman PRA, Blokland A, Bollen EPP et al. Phosphodiesterase inhibition
and modulation of corticostriatal and hippocampal circuits: clinical over-
view and translational considerations. Neurosci Biobehav Rev 2018;87:233-
54.
238. Kuroiwa M, Snyder GL, Shuto T et al. Phosphodiesterase 4 inhibition en-
hances the dopamine D1 receptor/PKA/DARPP-32 signaling cascade in
World Psychiatry 19:1 - February 2020 33
frontal cortex. Psychopharmacology 2012;219:1065-79.
239. Cohen SM, Tsien RW, Go DC et al. e impact of NMDA receptor hypo-
function on GABAergic neurons in the pathophysiology of schizophrenia.
Schizophr Res 2015;167:98-107.
240. Alherz F, Alherz M, Almusawi H. NMDAR hypofunction and somatostatin-
expressing GABAergic interneurons and receptors: a newly identied cor-
relation and its eects in schizophrenia. Schizophr Res Cogn 2017;8:1-6.
241. Singh SP, Singh V. Meta-analysis of the ecacy of adjunctive NMDA recep-
tor modulators in chronic schizophrenia. CNS Drugs 2011;25:859-85.
242. Weiser M, Heresco-Levy U, Davidson M et al. A multicenter, add-on ran-
domized controlled trial of low-dose d-serine for negative and cognitive
symptoms of schizophrenia. J Clin Psychiatry 2012;73:e728-34.
243. Iwata Y, Nakajima S, Suzuki T et al. Eects of glutamate positive modula-
tors on cognitive decits in schizophrenia: a systematic review and meta-
analysis of double-blind randomized controlled trials. Mol Psychiatry 2015;
20:1151-60.
244. Beck K, Javitt DC, Howes OD. Targeting glutamate to treat schizophre-
nia: lessons from recent clinical studies. Psychopharmacology 2016;233:
2425-28.
245. Go D. Bitopertin: the good news and bad news. JAMA Psychiatry 2014;
71:621-2.
246. D’Souza DC, Carson RE, Driesen N et al. Dose-related target occupancy and
eects on circuitry, behavior, and neuroplasticity of the glycine transporter-1
inhibitor PF-03463275 in healthy and schizophrenia subjects. Biol Psychiatry
2018;84:413-21.
247. Balu DT, Coyle JT. e NMDA receptor “glycine modulatory site” in schizo-
phrenia: D-serine, glycine, and beyond. Curr Opin Pharmacol 2015;20:109-15.
248. Zheng W, Cai DB, Yang XH et al. Adjunctive celecoxib for schizophrenia: a
meta-analysis of randomized, double-blind, placebo-controlled trials. J Psy-
chiatr Res 2017;92:139-46.
249. Zakrocka I, Targowska-Duda KM, Wnorowski A et al. Inuence of cyclooxy-
genase-2 inhibitors on kynurenic acid production in rat brain in vitro. Neu-
rotox Res 2019;35:244-54.
250. Anticevic A, Cole MW, Repovs G et al. Connectivity, pharmacology, and
computation: toward a mechanistic understanding of neural system dys-
function in schizophrenia. Front Psychiatry 2013;4:169.
251. Ellaithy A, Younkin J, González-Maeso J et al. Positive allosteric modulators
of metabotropic glutamate 2 receptors in schizophrenia treatment. Trends
Neurosci 2015;38:506-16.
252. Krystal JH, Abi-Saab W, Perry E et al. Preliminary evidence of attenuation
of the disruptive eects of the NMDA glutamate receptor antagonist, keta-
mine, on working memory by pretreatment with the group II metabotropic
glutamate receptor agonist, LY354740, in healthy human subjects. Psycho-
pharmacology 2005;179:303-9.
253. Uno Y, Coyle JT. Glutamate hypothesis in schizophrenia. Psychiatry Clin
Neurosci 2019;73:204-15.
254. Pillinger T, Rogdaki M, Mccutcheon RA et al. Altered glutamatergic response
and functional connectivity in treatment resistant schizophrenia: the eect of
riluzole and therapeutic implications. Psychopharmacology 2019;236:1985-
97.
255. Farokhnia M, Sabzabadi M, Pourmahmoud H et al. A double-blind, placebo
controlled, randomized trial of riluzole as an adjunct to risperidone for treat-
ment of negative symptoms in patients with chronic schizophrenia. Psycho-
pharmacology 2014;231:533-42.
256. Anand A, Charney DS, Oren DA et al. Attenuation of the neuropsychiatric
eects of ketamine with lamotrigine: support for hyperglutamatergic eects
of N-methyl-D-aspartate receptor antagonists. Arch Gen Psychiatry 2000;57:
270-6.
257. Tiihonen J, Wahlbeck K, Kiviniemi V. e ecacy of lamotr igine in clozapine-
resistant schizophrenia: a systematic review and meta-analysis. Schizophr
Res 2009;109:10-4.
258. Mouchlianitis E, McCutcheon R, Howes OD. Brain-imaging studies of treat-
ment-resistant schizophrenia: a systematic review. Lancet Psychiatry 2016;3:
451-63.
259. Egerton A, Broberg BV, Van Haren N et al. Response to initial antipsychotic
treatment in rst episode psychosis is related to anterior cingulate gluta-
mate levels: a multicentre 1H-MRS study (OPTiMiSE). Mol Psychiatry 2018;23:
2145-55.
260. Demjaha A, Egerton A, Murray RM et al. Antipsychotic treatment resistance
in schizophrenia associated with elevated glutamate levels but normal do-
pamine function. Biol Psychiatry 2014;75:e11-3.
261. Mouchlianitis E, Bloomeld MAP, Law V et al. Treatment-resistant schizo-
phrenia patients show elevated anterior cingulate cortex glutamate com-
pared to treatment-responsive. Schizophr Bull 2016;42:744-52.
262. Kennedy JA, Israel O, Frenkel A et al. Super-resolution in PET imaging. IEEE
Trans Med Imaging 2006;25:137-47.
263. Song T-A, Chowdhury SR, Yang F et al. Super-resolution PET imaging using
convolutional neural networks. Submitted for publication.
264. McCutcheon R, Pillinger T, Mizuno Y et al. e ecacy and heterogeneity
of antipsychotic response in schizophrenia: a meta-analysis. Mol Psychiatry
(in press).
265. Javitt DC, Carter CS, Krystal JH et al. Utility of imaging-based biomarkers for
glutamate-targeted drug development in psychotic disorders. JAMA Psy-
chiatry 2018;75:11-9.
266. van den Heuvel MP, Scholtens LH, Kahn RS. Multiscale neuroscience of
psychiatric disorders. Biol Psychiatry 2019;86:512-22.
267. Breakspear M. Dynamic models of large-scale brain activity. Nat Neurosci
2017;20:340-52.
DOI:10.1002/wps.20693
... Rs53576 polymorphism in the OXTR gene is found to be associated with striatal dopamine transporter availability in healthy subjects, and G carriers of this polymorphism could be more susceptible to environmental influences since the negative association of plasma oxytocin level with the striatal dopamine transporter availability was only observed in the G allele carriers of OXTR rs53576 [164]. Dysfunction of the dopamine system is a key factor in the pathophysiology of depression and schizophrenia [165][166][167], and environmental factors may play a significant role in dopaminergic dysregulation [167]. Exposure to diesel exhaust has been found to activate dopaminergic neurotoxicity in rats [168]. ...
... Rs53576 polymorphism in the OXTR gene is found to be associated with striatal dopamine transporter availability in healthy subjects, and G carriers of this polymorphism could be more susceptible to environmental influences since the negative association of plasma oxytocin level with the striatal dopamine transporter availability was only observed in the G allele carriers of OXTR rs53576 [164]. Dysfunction of the dopamine system is a key factor in the pathophysiology of depression and schizophrenia [165][166][167], and environmental factors may play a significant role in dopaminergic dysregulation [167]. Exposure to diesel exhaust has been found to activate dopaminergic neurotoxicity in rats [168]. ...
Article
Full-text available
Air pollution is one of the greatest environmental risks to health, with 99% of the world’s population living where the World Health Organization’s air quality guidelines were not met. In addition to the respiratory and cardiovascular systems, the brain is another potential target of air pollution. Population- and experiment-based studies have shown that air pollution may affect mental health through direct or indirect biological pathways. The evidence for mental hazards associated with air pollution has been well documented. However, previous reviews mainly focused on epidemiological associations of air pollution with some specific mental disorders or possible biological mechanisms. A systematic review is absent for early effect biomarkers for characterizing mental health hazards associated with ambient air pollution, which can be used for early warning of related mental disorders and identifying susceptible populations at high risk. This review summarizes possible biomarkers involved in oxidative stress, inflammation, and epigenetic changes linking air pollution and mental disorders, as well as genetic susceptibility biomarkers. These biomarkers may provide a better understanding of air pollution’s adverse effects on mental disorders and provide future research direction in this arena.
... These findings suggest that the SCZ-associated variant, rs4129585, likely lies within a TFBS that elicits its regulatory function in glutamatergic neurons and regions of the brain that have been implicated in neurodevelopment and SCZ pathology. [54][55][56][57][58][59][60][61][62][63] iPSC-derived NPCs exhibit transcriptomic changes associated with an SCZ risk allele With evidence that rs4129585 lies within a functional regulatory element implicated in neurobiology, we obtained human-derived iPSCs, used CRISPR-mediated genome editing to engineer the rs4129585 risk and non-risk alleles into otherwise isogenic cell lines, and differentiated these lines into NPCs. NPCs were selected as the differentiation endpoint because they represent an early stage in neuronal differentiation, and SCZ is considered a neurodevelopmental disorder. ...
... NPCs were selected as the differentiation endpoint because they represent an early stage in neuronal differentiation, and SCZ is considered a neurodevelopmental disorder. 60,61,64,65 Immunostaining and RT-qPCR with NPC-and iPSC-specific markers showed the former to be upregulated and the later downregulated ( Figures S2-S5), thus confirming successful differentiation from iPSCs to NPCs. We then performed RNA-seq to identify the transcriptional changes elicited by the altered variants. ...
Article
Full-text available
Recent collaborative genome-wide association studies (GWAS) have identified >200 independent loci contributing to risk for schizophrenia (SCZ). The genes closest to these loci have diverse functions, supporting the potential involvement of multiple relevant biological processes, yet there is no direct evidence that individual variants are functional or directly linked to specific genes. Nevertheless, overlap with certain epigenetic marks suggest that most GWAS-implicated variants are regulatory. Based on the strength of association with SCZ and the presence of regulatory epigenetic marks, we chose one such variant near TSNARE1 and ADGRB1, rs4129585, to test for functional potential and assay differences that may drive the pathogenicity of the risk allele. We observed that the variant-containing sequence drives reporter expression in relevant neuronal populations in zebrafish. Next, we introduced each allele into human induced pluripotent cells and differentiated four isogenic clones homozygous for the risk allele and five clones homozygous for the non-risk allele into neural progenitor cells. Employing RNA sequencing, we found that the two alleles yield significant transcriptional differences in the expression of 109 genes at a false discovery rate (FDR) of <0.05 and 259 genes at a FDR of <0.1. We demonstrate that these genes are highly interconnected in pathways enriched for synaptic proteins, axon guidance, and regulation of synapse assembly. Exploration of genes near rs4129585 suggests that this variant does not regulate TSNARE1 transcripts, as previously thought, but may regulate the neighboring ADGRB1, a regulator of synaptogenesis. Our results suggest that rs4129585 is a functional common variant that functions in specific pathways likely involved in SCZ risk.
... Neurotransmitter alterations are implicated in the pathophysiology of schizophrenia (43,44). To examine whether schizophrenia epicenter associated with neurotransmitter expressions, we investigated the spatial relationship between schizophrenia epicenter t map and 19 neurotransmitter receptor/transporter density maps (including dopamine, norepinephrine, serotonin, acetylcholine, glutamate, γaminobutyric acid (GABA), histamine, cannabinoid, and opioid), across nine different neurotransmitter systems (Extended Fig. 5A). ...
Article
Schizophrenia lacks a clear definition at the neuroanatomical level, capturing the sites of origin and progress of this disorder. Using a network-theory approach called epicenter mapping on cross-sectional magnetic resonance imaging from 1124 individuals with schizophrenia, we identified the most likely "source of origin" of the structural pathology. Our results suggest that the Broca's area and adjacent frontoinsular cortex may be the epicenters of neuroanatomical pathophysiology in schizophrenia. These epicenters can predict an individual's response to treatment for psychosis. In addition, cross-diagnostic similarities based on epicenter mapping over of 4000 individuals diagnosed with neurological, neurodevelopmental, or psychiatric disorders appear to be limited. When present, these similarities are restricted to bipolar disorder, major depressive disorder, and obsessive-compulsive disorder. We provide a comprehensive framework linking schizophrenia-specific epicenters to multiple levels of neurobiology, including cognitive processes, neurotransmitter receptors and transporters, and human brain gene expression. Epicenter mapping may be a reliable tool for identifying the potential onset sites of neural patho-physiology in schizophrenia.
... The GRINS project endeavors to comprehensively characterize the neurophysiological abnormities of SCZ and other psychiatric illnesses, primarily focuses on the thalamocortical system. Dysfunction in the thalamus and its cortical connections is believed to be related to the dysregulation of dopamine and glutamate neurotransmission, both of which are implicated in the pathophysiology of SCZ [56]. The thalamocortical system plays a pivotal role in the integration of sensory information and in the processing of higher cognitive functions such as perception, attention, and memory. ...
Article
Full-text available
Background Objective and quantifiable markers are crucial for developing novel therapeutics for mental disorders by 1) stratifying clinically similar patients with different underlying neurobiological deficits and 2) objectively tracking disease trajectory and treatment response. Schizophrenia is often confounded with other psychiatric disorders, especially bipolar disorder, if based on cross-sectional symptoms. Awake and sleep EEG have shown promise in identifying neurophysiological differences as biomarkers for schizophrenia. However, most previous studies, while useful, were conducted in European and American populations, had small sample sizes, and utilized varying analytic methods, limiting comprehensive analyses or generalizability to diverse human populations. Furthermore, the extent to which wake and sleep neurophysiology metrics correlate with each other and with symptom severity or cognitive impairment remains unresolved. Moreover, how these neurophysiological markers compare across psychiatric conditions is not well characterized. The utility of biomarkers in clinical trials and practice would be significantly advanced by well-powered transdiagnostic studies. The Global Research Initiative on the Neurophysiology of Schizophrenia (GRINS) project aims to address these questions through a large, multi-center cohort study involving East Asian populations. To promote transparency and reproducibility, we describe the protocol for the GRINS project. Methods The research procedure consists of an initial screening interview followed by three subsequent sessions: an introductory interview, an evaluation visit, and an overnight neurophysiological recording session. Data from multiple domains, including demographic and clinical characteristics, behavioral performance (cognitive tasks, motor sequence tasks), and neurophysiological metrics (both awake and sleep electroencephalography), are collected by research groups specialized in each domain. Conclusion Pilot results from the GRINS project demonstrate the feasibility of this study protocol and highlight the importance of such research, as well as its potential to study a broader range of patients with psychiatric conditions. Through GRINS, we are generating a valuable dataset across multiple domains to identify neurophysiological markers of schizophrenia individually and in combination. By applying this protocol to related mental disorders often confounded with each other, we can gather information that offers insight into the neurophysiological characteristics and underlying mechanisms of these severe conditions, informing objective diagnosis, stratification for clinical research, and ultimately, the development of better-targeted treatment matching in the clinic.
... This finding is in line with previous studies that were conducted in different samples where rumination was a key symptom (e.g. schizophrenia, post-traumatic stress disorder, and even subclinical anxiety) [71][72][73], therefore, may demonstrate that aberrant glutamatergic function and/or dysfunction in glutamate neurotransmission as a cardinal feature of rumination and disorders where rumination is a transdiagnostic factor. Though further investigation is warranted, decreasing frontal GLU concentrations and/or increasing frontal GABA concentrations via dietary approaches may hold promise as alternative interventions for these disorders, given that adherence to low GLU diets (whole food diets that restrict consumption of free glutamate and aspartate that are mainly found as flavour-enhancing food additives and in some foods including soy sauce, fish sauces, and aged cheeses) and consumption of high dietary GABA (found naturally in tea, tomato, soybean, germinated rice, fermented foods, and could be produced by lactic acid bacteria fermentation) have already been shown to affect brain and behaviour (improved subjective stress and sleep-related outcomes and altered cognitive performance, blood oxygen level dependent (BOLD) response and functional connectivity) [18,74,75]. ...
Article
Full-text available
Objectives: Common mental disorders (CMD) are associated with impaired frontal excitatory/ inhibitory (E/I) balance and reduced grey matter volume (GMV). Larger GMV (in the areas that are implicated in CMD-pathology) and improved CMD-symptomatology have been observed in individuals who adhere to high quality diets. Moreover, preclinical studies have shown altered neurometabolites (primarily gamma-aminobutyric acid: GABA and glutamate: GLU) in relation to diet quality. However, neurochemical correlates of diet quality and how these neurobiological changes are associated with CMD and with its transdiagnostic factor, rumination, is unknown in humans. Therefore, in this study, we examined the associations between diet quality and frontal cortex neuro-chemistry and structure, as well as CMD and rumination in humans. Methods: Thirty adults were classified into high and low diet quality groups and underwent 1H-MRS to measure medial prefrontal cortex (mPFC) metabolite concentrations and volumetric imaging to measure GMV. Results: Low (vs High) diet quality group had reduced mPFC-GABA and elevated mPFC-GLU concentrations, as well as reduced right precentral gyrus (rPCG) GMV. However, CMD and rumination were not associated with diet quality. Notably, we observed a significant negative correlation between rumination and rPCG-GMV and a marginally significant association between rumination and mPFC-GLU concentrations. There was also a marginally significant association between mPFC-GLU concentrations and rPCG-GMV. Discussion: Adhering to unhealthy dietary patterns may be associated with compromised E/I balance, and this could affect GMV, and subsequently, rumination.
Article
Full-text available
Schizophrenia (SZ) is a complex neuropsychiatric disorder associated with severe cognitive dysfunction. Although research has mainly focused on forebrain abnormalities, emerging results support the involvement of the cerebellum in SZ physiopathology, particularly in Cognitive Impairment Associated with SZ (CIAS). Besides its role in motor learning and control, the cerebellum is implicated in cognition and emotion. Recent research suggests that structural and functional changes in the cerebellum are linked to deficits in various cognitive domains including attention, working memory, and decision-making. Moreover, cerebellar dysfunction is related to altered cerebellar circuit activities and connectivity with brain regions associated with cognitive processing. This review delves into the role of the cerebellum in CIAS. We initially consider the major forebrain alterations in CIAS, addressing impairments in neurotransmitter systems, synaptic plasticity, and connectivity. We then focus on recent findings showing that several mechanisms are also altered in the cerebellum and that cerebellar communication with the forebrain is impaired. This evidence implicates the cerebellum as a key component of circuits underpinning CIAS physiopathology. Further studies addressing cerebellar involvement in SZ and CIAS are warranted and might open new perspectives toward understanding the physiopathology and effective treatment of these disorders.
Article
Full-text available
Importance Schizophrenia is a common, severe mental illness that most clinicians will encounter regularly during their practice. This report provides an overview of the clinical characteristics, epidemiology, genetics, neuroscience, and psychopharmacology of schizophrenia to provide a basis to understand the disorder and its treatment. This educational review is integrated with a clinical case to highlight how recent research findings can inform clinical understanding. Observations The first theme considered is the role of early-life environmental and genetic risk factors in altering neurodevelopmental trajectories to predispose an individual to the disorder and leading to the development of prodromal symptoms. The second theme is the role of cortical excitatory-inhibitory imbalance in the development of the cognitive and negative symptoms of the disorder. The third theme considers the role of psychosocial stressors, psychological factors, and subcortical dopamine dysfunction in the onset of the positive symptoms of the disorder. The final theme considers the mechanisms underlying treatment for schizophrenia and common adverse effects of treatment. Conclusions and Relevance Schizophrenia has a complex presentation with a multifactorial cause. Nevertheless, advances in neuroscience have identified roles for key circuits, particularly involving frontal, temporal, and mesostriatal brain regions, in the development of positive, negative, and cognitive symptoms. Current pharmacological treatments operate using the same mechanism, blockade of dopamine D2 receptor, which contribute to their adverse effects. However, the circuit mechanisms discussed herein identify novel potential treatment targets that may be of particular benefit in symptom domains not well served by existing medications.
Article
Full-text available
Preclinical models of psychosis propose that hippocampal glutamatergic neuron hyperactivity drives increased striatal dopaminergic activity, which underlies the development of psychotic symptoms. The aim of this study was to examine the relationship between hippocampal glutamate and subcortical dopaminergic function in people at clinical high risk for psychosis, and to assess the association with the development of psychotic symptoms. ¹H-MRS was used to measure hippocampal glutamate concentrations, and ¹⁸F-DOPA PET was used to measure dopamine synthesis capacity in 70 subjects (51 people at clinical high risk for psychosis and 19 healthy controls). Clinical assessments were undertaken at baseline and follow-up (median 15 months). Striatal dopamine synthesis capacity predicted the worsening of psychotic symptoms at follow-up (r = 0.35; p < 0.05), but not transition to a psychotic disorder (p = 0.22), and was not significantly related to hippocampal glutamate concentration (p = 0.13). There were no differences in either glutamate (p = 0.5) or dopamine (p = 0.5) measures in the total patient group relative to controls. Striatal dopamine synthesis capacity at presentation predicts the subsequent worsening of sub-clinical total and psychotic symptoms, consistent with a role for dopamine in the development of psychotic symptoms, but is not strongly linked to hippocampal glutamate concentrations.
Article
Full-text available
There is a large body of research reporting high rates of psychotic disorders among many migrant and minority ethnic groups, particularly in Northern Europe. In the context of increasing migration and consequent cultural diversity in many places worldwide, these findings are a major social and public health concern. In this paper, we take stock of the current state of the art, reviewing evidence on variations in rates of psychoses and putative explanations, including relevant theories and models. We discuss in particular: a) the wide variation in reported rates of psychotic disorders by ethnic group, and b) the evidence implicating social risks to explain this variation, at ecological and individual levels. We go on to set out our proposed socio‐developmental model, that posits greater exposure to systemic social risks over the life course, particularly those involving threat, hostility and violence, to explain high rates of psychoses in some migrant and minority ethnic groups. Based on this analysis, the challenge of addressing this social and public health issue needs to be met at multiple levels, including social policy, community initiatives, and mental health service reform.
Article
Full-text available
The response to antipsychotic treatment in schizophrenia appears to vary, and as such it has been proposed that different subtypes of schizophrenia exist, defined by treatment-response. This has not been formally examined using meta-analysis. Randomised controlled trials comparing placebo and antipsychotics in acute treatment of schizophrenia listed in PubMed, EMBASE and PsycINFO from inception until 30 November 2018 were examined. Relative variability of symptomatic improvement in antipsychotic-treated individuals compared to placebo-treated individuals was quantified using coefficient of variation ratio (CVR). Mean difference in symptom change was quantified using Hedges’ g. In addition, individual patient data from two clinical trials was examined in terms of both the distribution of total symptom change, and the variability of individual symptoms and symptom factors. In total, 11,006 articles were identified. Sixty six met inclusion criteria, reporting on 17,202 patients. Compared with placebo, antipsychotic-treated patients demonstrated greater total symptom improvement (g = 0.47, p < 0.001) and reduced variability in symptomatic improvement for total (CVR = 0.86, p < 0.001), positive (CVR = 0.89, p < 0.001), and negative symptoms (CVR = 0.86, p = 0.001). Lower variability in antipsychotic-response relative to placebo was associated with studies published earlier (z = 3.98, p < 0.001), younger patients (z = 3.07, p = 0.002), higher dose treatments (z = −2.62, p = 0.009), and greater mean-difference in symptom-change (z = −5.70, p < 0.001). In the individual patient dataset (N = 522 patients), antipsychotic treated patients did not show significantly increased variability for any individual symptom, and there was no evidence of a bimodal distribution of response. Compared to placebo, antipsychotic treatment shows greater improvement and lower variability of change in total, positive and negative symptoms. This is contrary to the hypothesis that there is a subtype of antipsychotic non-responsive schizophrenia. Instead our findings, provide evidence for a relatively homogeneous effect of antipsychotic treatment in improving symptoms of schizophrenia.
Article
Full-text available
Background: It has been hypothesized that dopamine function in schizophrenia exhibits heterogeneity in excess of that seen in the general population. However, no previous study has systematically tested this hypothesis. Methods: We employed meta-analysis of variance to investigate interindividual variability of striatal dopaminergic function in patients with schizophrenia and in healthy control subjects. We included 65 studies that reported molecular imaging measures of dopamine synthesis or release capacities, dopamine D2/3 receptor (D2/3R) or dopamine transporter (DAT) availabilities, or synaptic dopamine levels in 983 patients and 968 control subjects. Variability differences were quantified using variability ratio (VR) and coefficient of variation ratio. Results: Interindividual variability of striatal D2/3R (VR = 1.26, p < .0001) and DAT (VR = 1.31, p = .01) availabilities and synaptic dopamine levels (VR = 1.38, p = .045) but not dopamine synthesis (VR = 1.12, p = .13) or release (VR = 1.08, p = .70) capacities were significantly greater in patients than in control subjects. Findings were robust to variability measure. Mean dopamine synthesis (g = 0.65, p = .004) and release (g = 0.66, p = .03) capacities, as well as synaptic levels (g = 0.78, p = .0006), were greater in patients overall, but mean synthesis capacity did not differ from that of control subjects in treatment-resistant patients (p > .3). Mean D2/3R (g = 0.17, p = .14) and DAT (g = -0.20, p = .28) availabilities did not differ between groups. Conclusions: Our findings demonstrate significant heterogeneity of striatal dopamine function in schizophrenia. They suggest that while elevated dopamine synthesis and release capacities may be core features of the disorder, altered D2/3R and DAT availabilities and synaptic dopamine levels may occur only in a subgroup of patients. This heterogeneity may contribute to variation in treatment response and side effects.
Article
Full-text available
Background: Schizophrenia is one of the most common, burdensome, and costly psychiatric disorders in adults worldwide. Antipsychotic drugs are its treatment of choice, but there is controversy about which agent should be used. We aimed to compare and rank antipsychotics by quantifying information from randomised controlled trials. Methods: We did a network meta-analysis of placebo-controlled and head-to-head randomised controlled trials and compared 32 antipsychotics. We searched Embase, MEDLINE, PsycINFO, PubMed, BIOSIS, Cochrane Central Register of Controlled Trials (CENTRAL), WHO International Clinical Trials Registry Platform, and ClinicalTrials.gov from database inception to Jan 8, 2019. Two authors independently selected studies and extracted data. We included randomised controlled trials in adults with acute symptoms of schizophrenia or related disorders. We excluded studies in patients with treatment resistance, first episode, predominant negative or depressive symptoms, concomitant medical illnesses, and relapse-prevention studies. Our primary outcome was change in overall symptoms measured with standardised rating scales. We also extracted data for eight efficacy and eight safety outcomes. Differences in the findings of the studies were explored in metaregressions and sensitivity analyses. Effect size measures were standardised mean differences, mean differences, or risk ratios with 95% credible intervals (CrIs). Confidence in the evidence was assessed using CINeMA (Confidence in Network Meta-Analysis). The study protocol is registered with PROSPERO, number CRD42014014919. Findings: We identified 54 417 citations and included 402 studies with data for 53 463 participants. Effect size estimates suggested all antipsychotics reduced overall symptoms more than placebo (although not statistically significant for six drugs), with standardised mean differences ranging from -0·89 (95% CrI -1·08 to -0·71) for clozapine to -0·03 (-0·59 to 0·52) for levomepromazine (40 815 participants). Standardised mean differences compared with placebo for reduction of positive symptoms (31 179 participants) varied from -0·69 (95% CrI -0·86 to -0·52) for amisulpride to -0·17 (-0·31 to -0·04) for brexpiprazole, for negative symptoms (32 015 participants) from -0·62 (-0·84 to -0·39; clozapine) to -0·10 (-0·45 to 0·25; flupentixol), for depressive symptoms (19 683 participants) from -0·90 (-1·36 to -0·44; sulpiride) to 0·04 (-0·39 to 0·47; flupentixol). Risk ratios compared with placebo for all-cause discontinuation (42 672 participants) ranged from 0·52 (0·12 to 0·95; clopenthixol) to 1·15 (0·36 to 1·47; pimozide), for sedation (30 770 participants) from 0·92 (0·17 to 2·03; pimozide) to 10·20 (4·72 to 29·41; zuclopenthixol), for use of antiparkinson medication (24 911 participants) from 0·46 (0·19 to 0·88; clozapine) to 6·14 (4·81 to 6·55; pimozide). Mean differences compared to placebo for weight gain (28 317 participants) ranged from -0·16 kg (-0·73 to 0·40; ziprasidone) to 3·21 kg (2·10 to 4·31; zotepine), for prolactin elevation (21 569 participants) from -77·05 ng/mL (-120·23 to -33·54; clozapine) to 48·51 ng/mL (43·52 to 53·51; paliperidone) and for QTc prolongation (15 467 participants) from -2·21 ms (-4·54 to 0·15; lurasidone) to 23·90 ms (20·56 to 27·33; sertindole). Conclusions for the primary outcome did not substantially change after adjusting for possible effect moderators or in sensitivity analyses (eg, when excluding placebo-controlled studies). The confidence in evidence was often low or very low. Interpretation: There are some efficacy differences between antipsychotics, but most of them are gradual rather than discrete. Differences in side-effects are more marked. These findings will aid clinicians in balancing risks versus benefits of those drugs available in their countries. They should consider the importance of each outcome, the patients' medical problems, and preferences. Funding: German Ministry of Education and Research and National Institute for Health Research.
Article
Positron emission tomography (PET) suffers from severe resolution limitations which reduce its quantitative accuracy. In this paper, we present a super-resolution (SR) imaging technique for PET based on convolutional neural networks (CNNs). To facilitate the resolution recovery process, we incorporate high-resolution (HR) anatomical information based on magnetic resonance (MR) imaging. We introduce the spatial location information of the input image patches as additional CNN inputs to accommodate the spatially-variant nature of the blur kernels in PET. We compared the performance of shallow (3-layer) and very deep (20-layer) CNNs with various combinations of the following inputs: low-resolution (LR) PET, radial locations, axial locations, and HR MR. To validate the CNN architectures, we performed both realistic simulation studies using the BrainWeb digital phantom and clinical studies using neuroimaging datasets. For both simulation and clinical studies, the LR PET images were based on the Siemens HR+ scanner. Two different scenarios were examined in simulation: one where the target HR image is the ground-truth phantom image and another where the target HR image is based on the Siemens HRRT scanner — a high-resolution dedicated brain PET scanner. The latter scenario was also examined using clinical neuroimaging datasets. A number of factors affected relative performance of the different CNN designs examined, including network depth, target image quality, and the resemblance between the target and anatomical images. In general, however, all deep CNNs outperformed classical penalized deconvolution and partial volume correction techniques by large margins both qualitatively (e.g., edge and contrast recovery) and quantitatively (as indicated by three metrics: peak signal-to-noise-ratio, structural similarity index, and contrast-to-noise ratio).
Article
One of the most statistically significant loci to result from large-scale GWAS of schizophrenia is 10q24.32. However, it is still unclear how this locus is involved in the pathoaetiology of schizophrenia. The hypothesis that presynaptic dopamine dysfunction underlies schizophrenia is one of the leading theories of the pathophysiology of the disorder. Supporting this, molecular imaging studies show evidence for elevated dopamine synthesis and release capacity. Thus, altered dopamine function could be a potential mechanism by which this genetic variant acts to increase the risk of schizophrenia. We therefore tested the hypothesis that the 10q24.32 region confers genetic risk for schizophrenia through an effect on striatal dopamine function. To this aim we investigated the in vivo relationship between a GWAS schizophrenia-associated SNP within this locus and dopamine synthesis capacity measured using [18F]-DOPA PET in healthy controls. 92 healthy volunteers underwent [18F]-DOPA PET scans to measure striatal dopamine synthesis capacity (indexed as Kicer) and were genotyped for the SNP rs7085104. We found a significant association between rs7085104 genotype and striatal Kicer. Our findings indicate that the mechanism mediating the 10q24.32 risk locus for schizophrenia could involve altered dopaminergic function. Future studies are needed to clarify the neurobiological pathway implicated in this association.
Article
Antipsychotic drugs are central to the treatment of schizophrenia and other psychotic disorders but are ineffective for some patients and associated with side-effects and nonadherence in others. We review the in vitro, pre-clinical, clinical and molecular imaging evidence on the mode of action of antipsychotics and their side-effects. This identifies the key role of striatal dopamine D2 receptor blockade for clinical response, but also for endocrine and motor side-effects, indicating a therapeutic window for D2 blockade. We consider how partial D2/3 receptor agonists fit within this framework, and the role of off-target effects of antipsychotics, particularly at serotonergic, histaminergic, cholinergic, and adrenergic receptors for efficacy and side-effects such as weight gain, sedation and dysphoria. We review the neurobiology of schizophrenia relevant to the mode of action of antipsychotics, and for the identification of new treatment targets. This shows elevated striatal dopamine synthesis and release capacity in dorsal regions of the striatum underlies the positive symptoms of psychosis and suggests reduced dopamine release in cortical regions contributes to cognitive and negative symptoms. Current drugs act downstream of the major dopamine abnormalities in schizophrenia, and potentially worsen cortical dopamine function. We consider new approaches including targeting dopamine synthesis and storage, autoreceptors, and trace amine receptors, and the cannabinoid, muscarinic, GABAergic and glutamatergic regulation of dopamine neurons, as well as post-synaptic modulation through phosphodiesterase inhibitors. Finally, we consider treatments for cognitive and negative symptoms such dopamine agonists, nicotinic agents and AMPA modulators before discussing immunological approaches which may be disease modifying.