Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2016 Jun 10.
Published in final edited form as: J Psychopharmacol. 2015 Jan 13;29(2):97–115. doi: 10.1177/0269881114563634

Glutamate and dopamine in schizophrenia: an update for the 21st century

Oliver Howes 1,2,*, Rob McCutcheon 1,2, James Stone 1,2
PMCID: PMC4902122  EMSID: EMS66264  PMID: 25586400

Abstract

The glutamate and dopamine hypotheses are leading theories of the pathoaetiology of schizophrenia. Both were initially based on indirect evidence from pharmacological studies supported by post-mortem findings, but have since been substantially advanced by new lines of evidence from in vivo imaging studies. This review provides an up- date on the latest findings on dopamine and glutamate abnormalities in schizophrenia, focusing on the in vivo neuroimaging studies in patients and clinical high risk groups, and considers their implications for understanding the biology and treatment of schizophrenia. These findings have refined both the dopamine and glutamate hypotheses, enabling greater anatomical and functional specificity, and have been complemented by preclinical evidence showing how the risk factors for schizophrenia impact on the dopamine and glutamate systems. The implications of this new evidence for understanding the development and treatment of schizophrenia are considered, and the gaps in current knowledge highlighted. Finally the evidence for an integrated model of the interactions between the glutamate and dopamine systems is reviewed, and future directions discussed.

Keywords: schizophrenia, psychosis, mechanisms, treatment, antipsychotic, imaging, etiology, PET, MR, dopamine, glutamate, NMDA, D2

Introduction

Schizophrenia is a common, severe mental illness (Jablensky et al., 1992; Jablensky, 2000), and a leading cause of adult disease burden (Whiteford et al., 2013). It has a lifetime prevalence of about 0.7%, with a peak age of onset in the early twenties in men, and three or four years later in women (Saha et al., 2005). The disorder is characterised by psychotic symptoms such as delusions and hallucinations, negative symptoms such as social withdrawal and amotivation, and cognitive impairments. In addition to the high morbidity and mortality rate associated with schizophrenia, the health and social care costs for the illness are substantial: total costs of care for individuals with schizophrenia in the UK amount to 16.7 billion EURO per annum, while for Europe the figure is 93.9 billion EURO (Gustavsson et al., 2011). Antipsychotic drugs are the mainstay of treatment for schizophrenia, but are inadequate in a substantial proportion of patients (Andrews, 2003). Better understanding of the neurobiology underlying the disorder is needed to improve the use of existing drugs and inform the development of new drugs.

Two of the most influential hypotheses concerning the neurobiology underlying the disorder involve dopamine and glutamate. Both hypotheses were proposed several decades ago, but new evidence, particularly from in vivo imaging studies and preclinical findings on the role of these neurotransmitters has refined understanding of the nature of dopamine and glutamate dysfunction in schizophrenia. The purpose of this review is to provide an overview of both hypotheses and an up-date on the recent findings relevant to them, focussing on the in vivo neuroimaging findings. Finally it considers the degree to which they may be integrated, the implications of the new findings for treatment, and the limitations of current evidence.

Dopamine

Dopamine: early evidence for its involvement in schizophrenia

The dopamine hypothesis of schizophrenia initially arose from several indirect sources of evidence. A key source was studies showing that administration of amphetamine and other compounds which increase extracellular concentrations of dopamine can induce psychotic symptoms similar to those seen in schizophrenia (see review (Lieberman et al., 1987)). This was supported by studies of drugs, such as reserpine and alpha-methyl-para-tyrosine that deplete dopamine levels (Carlsson et al., 1957), which showed that these drugs reduced psychotic symptoms (Campden-Main and Wegielski, 1955; Arnold and Freeman, 1956; Carlsson et al., 1973; Walinder and Skott, 1976).

Further evidence for dopamine’s involvement in schizophrenia came in the 1970s, with observations that the clinical effectiveness of antipsychotic drugs was directly related to their affinity for dopamine receptors (Seeman and Lee, 1975; Seeman et al., 1976; Creese et al., 1976). As a result of this discovery, the leading hypothesis at this time was that schizophrenia arises as a result of abnormalities in dopamine receptor density (Matthysse, 1973; Snyder, 1976).

Whilst these findings implicated dopamine in schizophrenia, this was in a rather general way and there were a number of limitations on the interpretation of the evidence. For example, both amphetamine and reserpine affect other brain monoamines as well as dopamine (see review (Davis et al., 1991)). Furthermore, at this time there was no clear indication of the locus of dopaminergic abnormality in the living brain.

Post-mortem studies are able to provide anatomical detail and biochemical specificity. Early post-mortem studies suggested that the neuropathological changes in schizophrenia included both an increase in striatal dopamine levels, and an increase in D2 receptor density (Owen et al., 1978; Mackay et al., 1982), but no change in dopamine transporter (DAT) densities (Pearce and Seeman, 1990).

More recent studies have provided greater detail regarding the nature of these pre and post-synpatic dopamine changes. For example, recent research has shown tyrosine hydroxylase, the rate-limiting enzyme involved in the synthesis of dopamine, is significantly increased in the substantia nigra of patients with schizophrenia compared to patients with depression and healthy controls (Howes et al., 2013a), indicating that there is increased capacity for production of dopamine in the midbrain origin of dopamine neurons as well as their striatal terminals. Other recent work has refined understanding of the nature of the D2 receptor changes. A study including 176 post mortem samples from patients with schizophrenia showed the expression of the presynaptic D2 autoreceptor was increased, while the expression of predominantly postsynaptic variants were decreased in the dorsolateral prefrontal cortex compared with controls (Kaalund et al., 2013). The hypothesis that D2 receptors are somehow altered in schizophrenia is supported by genetics’ findings that have shown a clear association between the DRD2 gene and schizophrenia. (Ripke et al., 2014).

The potential relevance of the D4 receptor was initially highlighted by post mortem work showing a sixfold increase in D4 receptor density in schizophrenia (Seeman et al., 1993). This finding appeared to bear particular relevance given clozapine’s unique clinical properties and its high level of binding to the D4 receptor not shared by other antipsychotics. Subsequent attempts at replication, however, were mixed (Reynolds and Mason, 1994; Murray et al., 1995).

Dopamine receptors form dimers with themselves and with other receptors in in vitro models. A number of studies have recently looked at the role that dimerisation of the dopamine receptor may play in schizophrenia. In a study of 15 schizophrenia patients Wang and colleagues found a 278% increase in expression of D2 dimers compared to controls, while D2 monomers were decreased to 69% of the control level (Wang et al., 2010). In addition there is preliminary evidence concerning D2 heteromers from a post-mortem study that found that D1-D2 heteromers were increased in schizophrenia in the globus pallidus, although the study included only 4 patients with schizophrenia (Perreault et al., 2010). This is complemented by earlier post-mortem findings that showed the inhibitory link between D1 and D2 receptors is reduced in patients with schizophrenia (Seeman et al., 1989). These findings highlight the potential impact of changes in dimerization, although, as it is not yet possible to directly measure dimerization in vivo, these finding await testing in vivo in patients.

In summary, post-mortem studies continue to identify abnormalities in both the presynaptic and post-synaptic dopaminergic system in schizophrenia. However, one potential limitation of post mortem studies is the difficulty in controlling for the confounding effects of antipsychotic medication, and it is plausible that the presynaptic and post-synaptic changes observed are predominantly iatrogenic. Another difficulty is linking the changes to the expression of symptoms and the development of the disorder, often many years before the patient died. In order to address these issues evidence from living patients is required. The following sections consider the latest in vivo neurochemical imaging evidence in patients.

Dopamine and schizophrenia: in vivo imaging evidence

Positron Emission Tomography (PET) and Single Photon Emission Computed Tomography (SPECT) imaging allows the in vivo quantification of many aspects of dopaminergic function in the brain, including dopamine synthesis, the degree of dopamine release in response to stimuli, and the availability of the post-synaptic dopaminergic receptors and transporters (Kim et al., 2013a). Over the last two decades or so, advances in PET/SPECT technology and their application has enabled the major aspects of the dopaminergic hypothesis of schizophrenia to be tested and refined (McGuire et al., 2008). Studies initially focussed on dopamine D2 receptors and the action of antipsychotic drugs, before examining other aspects of the dopaminergic system

Antipsychotic drugs and dopamine D2 receptors

All currently licensed antipsychotic drugs block D2 dopamine receptors. However, they also act at other receptors in the brain, including other dopamine receptor subtypes and those for serotonin, histamine, norepinephrine and acetylcholine (Stahl, 2013). It was initially far from clear which receptor mediated the clinical response to antipsychotic treatment. Molecular imaging studies with both SPECT (e.g.: (Brücke et al., 1991; Brücke et al., 1992; Pilowsky et al., 1993; Volk et al., 1994; Klemm et al., 1996)) and PET (e.g.: (Farde et al., 1988; Baron et al., 1989; Wolkin, 1989; Nordström et al., 1992; Nordstrom et al., 1993; Goyer et al., 1996; Kapur et al., 1996)) have extended the in vitro studies on dopamine receptors and antipsychotics from the 1970s in several crucial ways. First, they have demonstrated that all antipsychotic drugs cross the blood-brain-barrier. Second, they have shown that they block D2/3 striatal receptors in vivo at clinically effective doses. These data extended the in vitro findings to patients, and have provided the foundation for studies in which the relationship between D2 occupancy and clinical response could be established.

The relationship between D2 occupancy, clinical response, and side effects is not linear (see review Howes et al. 2009a)). Little response is seen when occupancy is below 50%, response increases from this point but the risk of extrapyramidal side effects increases when occupancy reaches around 75% (Nordstrom et al., 1993). These findings have been replicated in a double blind study of first-episode patients. This study found a threshold D2 occupancy of 65% was found to best separate responders from non-responders: at 65% receptor occupancy, 80% of responders were above the threshold whilst 67% of the non–responders lay below the 65% threshold (Kapur et al., 2000). Given the central role of dopamine receptors in the mode of action of antipsychotics, it is not surprising that considerable focus has been directed at determining if there are alterations in D2 receptors in vivo.

Dopamine D2 Receptor availability in schizophrenia

The action of dopamine on post-synaptic receptors constitutes the final stage in transmitting the dopaminergic neuronal impulse to post-synaptic neurons. As all clinically available antipsychotics block D2 receptors, over the last three decades many studies have investigated whether D2 receptor availability is altered in schizophrenia compared to control subjects.

It is important to note that the radioligands used to image dopamine receptors in vivo such as [11C]raclopride (Malmberg et al., 1993) and [123I](S)-(-)-3-iodo-2-hydroxy-6-methoxy-N-[(1-ethyl-2-pyrrolidinyl)methyl]benzamide ([123I]IBZM) (Videbaek et al., 2000) show significant D3 receptor binding in addition to their D2 binding. Thus it is theoretically possible that an elevation in one could be obscured by a reduction in another. [11C]-(+)-4-propyl-9-hydroxynaphthoxazine ([11C]-PHNO) is an agonist radiotracer that shows higher affinity for D3 receptors than D2. Studies in schizophrenia using this tracer have not shown any difference in binding from controls (Graff-Guerrero et al., 2009). This suggests that alterations in D3 receptor availability are not masking changes in D2 receptors, although the development of selective D2 tracers would be useful to definitively establish this.

There have been at least twenty-two published studies, but, after an initial positive finding, subsequent results have been inconsistent. The most recent meta-analysis (Howes et al., 2012b) found increased D2/3 receptor density in patients with schizophrenia, but the effect size was small (d=0.26, p=0.049). However the picture is complicated by large heterogeneity between studies. In the subgroup analysis of studies that exclusively considered antipsychotic-naïve patients no significant differences from controls was found, whereas significant differences were found in the studies that included patients who had received antipsychotic treatment. Other sources of heterogeneity include different radiotracers used and differences in duration of illness. In addition, elevated baseline dopamine levels in schizophrenic patients could potentially mask differences in receptor density. Nevertheless the most likely implications are that there is not a major elevation in D2 receptor availability in drug naïve patients and that antipsychotic treatment leads to D2 receptor upregulation in some patients. The potential effect of antipsychotics on D2 receptors has been tested directly in a PET imaging study comparing D2 binding potentials between a group of treatment naïve patients with a group that had been treated with antipsychotics long term that underwent temporary antipsychotic withdrawal (Silvestri et al., 2000). Significantly increased D2 receptor availability was found in the group that had experienced long term antipsychotic treatment, consistent with the idea that there is D2 upregulation with antipsychotic treatment.

Changes in D1 receptor densities have been less thoroughly investigated and are complicated by variation in the populations studied and the radiotracers employed. Two studies looking at chronic, medicated patients have shown widespread cortical and striatal reductions in D1 receptor densities (Hirvonen et al., 2006; Kosaka et al., 2010). One study of antipsychotic naïve patients and one of naïve and previously treated patients showed increased densities of D1 receptors in the prefrontal cortex in patients (Abi-Dargham et al., 2002; Abi-Dargham et al., 2012). Conversely another study of antipsychotic naïve and medication free patients showed a decrease in prefrontal cortex densities (Okubo et al., 1997), while a further study of purely naïve patients showed no significant differences (Karlsson et al., 2002). Interpretation of the studies is further complicated by the discovery that the radiotracers used (NNC 112 and SCH 23390) also bind to the 5HT2A receptor (Ekelund et al., 2007; Catafau et al., 2010),. This serotonergic receptor has been implicated in the aetiology of schizophrenia, and its involvement has been further supported by a recent meta-analysis of post-mortem findings (Selvaraj et al., 2014). Further studies using more specific radiotracers are thus needed to determine the nature of D1 changes in schizophrenia.

Fewer studies have investigated D2/3 receptor density in extrastriatal brain regions. A recent meta-analysis did not find any significant differences for receptor densities in the thalamus, temporal cortex or substantia nigra; while in other brain regions where it was not possible to conduct a meta-analysis findings were either statistically insignificant or have not been replicated (Kambeitz et al., 2014).

Finally, the D2 receptor may exist in two intraconvertible (high and low) affinity states for agonist binding and it has been proposed that the balance between these two states is altered in schizophrenia (Seeman et al., 2006). An initial study with [11C]-(+)-PHNO ([11C]-(+)-4-propyl-9-hydroxynaphthoxazine), a radiotracer selective for the high affinity state as well as D3 receptors, suggested that no alteration is apparent in the absence of dopamine depletion (Graff-Guerrero et al., 2009). However further work has suggested that the radiotracer used may be unable to distinguish between the two receptor states (Seeman, 2012), indicating that this issue remains unresolved.

Dopamine transporter levels

The dopamine active transporter (DAT) regulates dopaminergic transmission by removing dopamine from the synaptic cleft. The DAT is solely present in dopamine synthesising axons (Ciliax and Heilman, 1995), and because of this specificity in location it is a useful measure of the integrity of dopaminergic neurons. Early neurodegenerative theories of schizophrenia proposed that a change in the density of striatal dopamine terminals was fundamental to the disorder; one strand of this hypothesis was that hyperdopaminergic activity could itself be neurotoxic and lead to loss of dopaminergic neurons in the striatum (Lieberman et al., 1990). Imaging of the DAT has allowed this hypothesis to be tested.

There have been thirteen PET or SPECT imaging studies that have addressed DAT density in schizophrenia. There is a large degree of heterogeneity between the studies with a number of inconsistent findings. Recent meta-analyses (Howes et al., 2012b; Chen et al., 2013) found no evidence for a difference in striatal DAT receptor densities in patients with schizophrenia, and this finding has been replicated in a subsequent study of almost fifty antipsychotic-naïve patients (Chen et al., 2013). This essentially negative result suggests that the dopaminergic abnormalities observed in schizophrenia are not secondary to abnormalities in the DAT, nor due to differences in the density of pre-synaptic dopaminergic neurons.

Dopamine synthesis and release

In dopamine neurons radiolabelled L-DOPA is converted to dopamine and trapped in dopaminergic nerve terminals. The extent of radiolabelled L-DOPA uptake measured using PET thus provides an index of presynaptic dopamine synthesis capacity and the availability of dopamine for release from presynaptic terminals (see review by (Kumakura and Cumming, 2009)), and shows good test retest reliability in dopaminergic regions (Egerton et al., 2010a). Meta-analysis of the studies using this technique shows a large elevation in dopamine synthesis capacity in schizophrenia compared with matched controls, with an effect size of 0.8 (Howes et al., 2012b). With the addition of three studies published after this meta-analysis, there have now been fourteen PET studies of dopamine synthesis capacity in patients with schizophrenia, summarised in table 1 (adapted from (Howes et al., 2012b)).

Table 1.

The PET imaging studies of striatal dopamine synthesis capacity in schizophrenia

Study Medication status MN/MF/M/HC Phenotype Radiotracer Findings (standard effect size)
Dao-Castellana et al.1997 2/4/0/7 Chronic [18F]-DOPA - (0.35)
Demjaha et al. 2012d 0/0/12/12 Chronic [18F]-DOPA ↑ (1.12)
Elkashef et al. 2000 0/9/10/13 Chronic [18F]-DOPA - (-0.13)
Reith et al.1994 4/1/0/13 Chronic [18F]-DOPA ↑ (1.52)
Hietala et al.1995a 7/0/0/8 FEP [18F]-DOPA - (0.9)
Hietala et al.1999 10/0/0/13 FEP [18F]-DOPA ↑ (1.02)
Howes et al. 2009 4/3/0/12 FEP/Chronic [18F]-DOPA ↑ (1.18)
Howes et al. 2013ae 5/8/16/29 Chronic [18F]-DOPA ↑ (1.14)
Kumakara et al. 2007b 3/5/0/15 Chronic [18F]-DOPA - (0.10)
Lindstrom et al. 1999 10/2/0/10 FEP/Chronic [11C]-DOPA ↑ (1.01)
McGowan et al. 2004 0/0/16/12 Chronic [18F]-DOPA ↑ (1.55)
Meyer-Lindenberg et al. 2002 0/6/0/6 Chronic [18F]-DOPA ↑ (1.82)
Nozaki et al. 2009c 14/4/0/20 FEP/Chronic [11C]-DOPA - (0.13)
Shotbolt et al. 2011 0/0/6/20 Chronic [18F]-DOPA - (NA)

↑ = significantly higher in patient group; ↓ = significantly lower level in patient group; - = no significant difference

[11C]DOPA= L-[β-11C]Dihydroxyphenylalanine; [18F]-DOPA- 6-[18F]fluoro-L-Dihydroxyphenylalanine

FEP- first episode psychosis; MN- antipsychotic naïve; HC – healthy control; M – currently taking antipsychotic medication; MF- antipsychotic free

a

Found significant difference between patients and controls found for the putamen but not striatum as a whole.

b

Found significantly increased [18F]-DOPA turnover in patients compared to controls.

c

Found significantly increased uptake in left caudate nucleus but not striatum as a whole.

d

Synthesis capacity significantly increased in the 12 responders to antipsychotic treatment but not in the treatment resistant patients

e

Includes 14 patients and 12 controls from McGowan et al. 2004.

Dopaminergic neurotransmission requires the release of dopamine from presynaptic terminals. This can be indexed in vivo using molecular imaging with radiotracers which bind to dopamine D2 receptors, such as [11C]raclopride or [123I]IBZM, as these radiotracers compete with dopamine to bind to dopamine D2 receptors. Decreased radiotracer binding following a pharmacological challenge that releases dopamine thus reflects increased extracellular dopamine (Egerton et al., 2010b). All six investigations that have used this approach in schizophrenia have found evidence of significantly increased dopamine release in patients compared to control subjects (Laruelle et al., 1999; Abi-Dargham et al., 2009; Pogarell et al., 2012) (see table 2). The extent of radiotracer displacement was approximately doubled in schizophrenic patients, and the degree of displacement correlated with the degree of worsening of psychotic symptoms. These findings have been recently been extended by using a social stress task in the place of the pharmacological challenge, and showing increased dopamine release to stress in schizophrenia (Mizrahi et al., 2012).

Table 2.

The PET imaging studies of dopamine release/baseline occupancy in schizophrenia

Study Medication status MN/MF/M/HC Phenotype Radiotracer Study Type Findings (standard effect size)
Abi-Dargham et al. 1998a 2/13/0/15 13 Chronic 2 FEP [123I]IBZM Amphetamine Induced ↑ (0.79)
Abi-Dargham et al. 2009c 6/0/0/8 FEP [123I]IBZM Amphetamine Induced ↑ (1.32)
Breier et al. 1997 4/7/0/12 Chronic [11C] raclopride Amphetamine Induced ↑ (0.88)
Laruelle et al. 1996 0/15/0/15 Chronic [123I]IBZM Amphetamine Induced ↑ (0.96)
Laruelle et al. 1999b 7/27/0/36 27 Chronic 7 FEP [123I]IBZM Amphetamine Induced ↑ (0.91)
Mizrahi et al. 2012 10/0/0/12 FEP [11C]-+-PHNO MIST Induced ↑ (1.37)
Pogarell et al. 2012 0/8/0/7 FEP [123I]IBZM Amphetamine Induced ↑ (1.19)
Abi-Dargham et al 2000 8/10/0/18 10 chronic 8 FEP [123I]IBZM AMPT depletion ↑ (1.08)
Kegeles et al. 2010d 6/12/0/18 FEP/Chronic [11C] raclopride AMPT depletion - (0.63)

↑ = significantly higher in patient group; ↓ = significantly lower level in patient group; - = no significant difference

[123I]IBZM - [123I](S)-(-)-3-iodo-2-hydroxy-6-methoxy-N- [(1-ethyl-2-pyrrolidinyl)methyl]benzamide; [11C]-+-PHNO - [11C]-(+)-4-propyl-9-hydroxynaphthoxazine; MIST – Montreal imaging stress task

FEP- first episode psychosis; MN- antipsychotic naïve; HC – healthy control; M – currently taking antipsychotic medication; MF- antipsychotic free

a

Abi-dargham et al.report a larger effect size of 1.06 as only the control group SD isincluded in their calculation (as opposed to both patient and control SDs)

b

Includes 30pts from Laruelle et al. 1996 and Abi-Dargham et al.1998

c

Includes patients from Abi-dargham et al. 2000

d

Significant change only found for associative striatum not striatum overall

An alternative method to index synaptic dopamine levels involves the use of alpha-methyl-p-tyrosine, a tyrosine hydroxylase inhibitor. This depletes intrasynaptic dopamine and therefore allows the level of baseline synaptic occupancy to be estimated from the subsequent increase in radiotracer binding. Two studies using this technique have shown increased baseline occupancy of D2 receptors in schizophrenia, indicating that extracellular dopamine concentrations are also increased at baseline in schizophrenia (Kegeles et al., 2010) (see table 2).

Presynaptic dopamine dysfunction: state or trait marker?

Together, these studies provide compelling evidence that presynaptic dopamine availability and dopamine release are increased in schizophrenia. This raises the question is increased presynaptic dopamine a state marker of psychosis or a trait marker related to risk of schizophrenia?

In the studies where patients were acutely psychotic at the time of investigation, elevated dopamine synthesis capacity has been consistently detected (Hietala et al., 1995; Hietala et al., 1999; Howes et al., 2009b), whereas elevated dopamine synthesis capacity has been less consistently detected in studies of chronic patients, with some studies not finding significant differences from controls (Dao-Castellana et al., 1997; Elkashef et al., 2000; Shotbolt et al., 2011) whilst others have detected significant elevations (Reith et al., 1994; Meyer-Lindenberg et al., 2002; Mcgowan et al., 2004; Howes et al., 2013a). Similarly patients experiencing an acute relapse show evidence of significantly greater dopamine release than controls, and although patients in remission show greater dopamine release than controls in absolute terms, this was not statistically significant (Laruelle et al., 1999).

In recent years a significant body of work has developed examining the prodromal phase of schizophrenia (see table 3). Howes et al. initially showed that dopamine synthesis capacity is increased in individuals with prodromal symptoms of schizophrenia (Howes et al., 2009b), now replicated in an independent cohort (Egerton et al., 2013), while later work showed that the increase in capacity was specific to individuals that went on to develop a psychotic illness (Howes et al., 2011b). Finally, the studies in the prodrome to schizophrenia provide evidence that dopamine synthesis capacity increases further with the onset of acute psychosis (Howes et al., 2011a).

Table 3.

The PET studies of dopaminergic function in individuals at increased risk of schizophrenia

Study Medication status MN/MF/M/HC Phenotype Radiotracer Study type Findings (standard effect size)
Bloemen et al. 2013b 14/0/0/15 UHR [123I]IBZM Dopamine depletion -
Howes et al. 2009 23/0/1/12 UHR [18F]-DOPA Synthesis capacity ↑ (0.75)
Howes et al. 2011ba 30/0/0/29 UHR [18F]-DOPA Synthesis capacity ↑ (1.18)
Mizrahi et al. 2013 12/0/0/12 UHR [11C]-+-PHNO MIST induced dopamine release ↑ (0.98)
Suridjan et al. 2013 12/0/0/12 UHR [11C]-+-PHNO D2high/D3 receptor availability -
Abi-dargham et al. 2004 13/0/0/13 Schizotypal [123I]IBZM Amphetamine induced dopamine release ↑ (0.93)
Howes et al. 2012 16/0/0/16 Auditory hallucinations (otherwise healthy) [18F]-DOPA Synthesis capacity -
Huttunen et al. 2008 17/0/0/17 First degree relatives of patients with schizophrenia [18F]-DOPA Synthesis capacity
Shotbolt et al. 2011 6/0/0/20 Twins of patients with schizophrenia [18F]-DOPA Synthesis capacity -
Soliman et al. 2008c 16/0/0/10 “Potential schizotype” [11C] raclopride MIST induced dopamine release

↑ = significantly higher in patient group; ↓ = significantly lower level in patient group; - = no significant difference

[123I]IBZM - [123I](S)-(-)-3-iodo-2-hydroxy-6-methoxy-N- [(1-ethyl-2-pyrrolidinyl)methyl]benzamide; [11C]-+-PHNO - [11C]-(+)-4-propyl-9-hydroxynaphthoxazine; [18F]-DOPA- 6-[18F]fluoro-L-Dihydroxyphenylalanine; MIST – Montreal Imaging Stress Task; D1, D2, D3= Dopamine receptor sub-type 1, 2 and 3 respectively

MN- antipsychotic naïve; HC – healthy control; M – currently taking antipsychotic medication; MF- antipsychotic free; UHR – Ultra high risk for psychosis

a

Synthesis capacity increased only in subgroup that transitioned to psychosis (n=9)

b

No significant differences between controls and UHR but correlation between receptor occupancy by dopamine and positive symptoms

c

MIST leads to significant decrease in tracer binding in subgroup characterized as ‘potential schizotype’ on the basis that subjects scored >1.95 SD on the negative subscale of Chapman schizotypy questionnaire.

Taken together these findings suggest there is at least a component of presynaptic dysfunction that changes and is related to the acute psychotic state. However, this does not exclude the possibility that a further component is related to risk.

In order to determine whether increased synthesis capacity is a marker for risk of schizophrenia two studies have examined individuals with an increased genetic risk of schizophrenia. In a study of seventeen first degree relatives of individuals with schizophrenia Huttenen et al. (Huttunen et al., 2008) found significantly increased dopamine synthesis capacity compared to controls. However, Shotbolt et al. did not find altered dopamine synthesis capacity in twins discordant for schizophrenia (Shotbolt et al., 2011). As the twins in this study were predominantly dizygotic it remains possible that they had not inherited the genetic risk, and further studies are needed to determine if genetic risk is associated with dopamine dysregulation per se.

A contrasting approach has been to see whether increased synthesis capacity is present in individuals who experience psychotic-like symptoms but do not have a diagnosis of schizophrenia. Early research showed that psychotic symptoms secondary to temporal lobe epilepsy were associated with increased striatal dopamine synthesis capacity to a similar degree to that seen in schizophrenia (Reith et al., 1994). This has also been shown to be the case in individuals with schizotypal personality disorder (Abi-Dargham et al., 2004), while this has not found in healthy individuals who experience persistent sub-clinical auditory hallucinations but have never developed schizophrenia or functional impairment (Howes et al., 2013b), suggesting presynaptic dopamine dysfunction is linked to clinical disorder rather than being a marker of psychotic-like experiences per se.

In summary, evidence from a number of studies indicates that presynaptic dopamine dysfunction has at least a state component, but the evidence that it is a trait marker is less clear cut. Research using healthy twin pairs has found evidence that environmental factors explain a substantial proportion of variation in normal presynaptic dopamine function (Stokes et al., 2013). This fits with recent models that stress and other risk factors for psychosis impact on a vulnerable dopamine system to dysregulate it and lead to psychosis (Howes and Murray, 2014).

Linking dopamine to symptoms and environmental risk factors

The evidence reviewed above indicates there is a presynaptic hyperdopaminergic abnormality in schizophrenia and that antipsychotics act by blocking D2 receptors to treat psychotic symptoms. In and of themselves however, these findings do not explain how a biochemical abnormality can account for the phenomenology of the psychotic experience. The aberrant salience model (Kapur, 2003; Heinz and Schlagenhauf, 2010; Winton-Brown et al., 2014)seeks to provide this link.

Animal studies provide support for dopamine’s role as a mediator in assigning motivational salience to internal or external stimuli, and thereby determining which stimuli grab attention and drive behaviour (Bromberg-Martin et al., 2010). It is proposed that in schizophrenia, the elevation in presynaptic dopamine leads to its release in the absence of appropriate stimuli (Winton-Brown et al., 2014). Dysregulated release is thought to lead to the attribution of salience to irrelevant stimuli simply by virtue of the stimuli’s temporal association with the dopaminergic signalling. This aberrant attribution of salience is thought to account for psychotic phenomena such as ideas of reference, and patients’ accounts of the prodromal phase of psychosis, where everyday occurrences are imbued with a sense of inexplicable significance (Fusar-Poli et al., 2008). There is certainly evidence that untreated and drug naïve patients with schizophrenia show alterations in processing motivationally salient stimuli (Juckel et al., 2006; Schlagenhauf et al., 2009; Nielsen et al., 2012). In addition there is also evidence that patients with current delusions and people with prodromal-type symptoms who are at risk of psychosis show greater assignment of importance to irrelevant stimuli, supporting the aberrant salience hypothesis (Roiser et al., 2009; Roiser et al., 2013). This hypothesis has been integrated with cognitive models (Garety et al., 2001) to explain the paranoid flavour of many delusional experiences by proposing that social adversity leads to the development of paranoid biases, which adds a persecutory colour to the misattribution of salience (Howes and Murray, 2014). The subsequent crystallisation of this experience into a delusion can be viewed as an individual’s attempt to construct a coherent narrative that accounts for their experiences, and influenced by their personal and cultural background.

The combination of this aberrant salience model with the dopaminergic abnormalities described above, and evidence on the impact of environmental and neurodevelopmental risk factors has contributed to the dopamine hypothesis being revised (Howes and Kapur, 2009; Howes and Murray, 2014). However this is predominantly a model of psychosis in schizophrenia.

Cognitive and negative symptoms of schizophrenia are responsible for a major proportion of the disability associated with the disorder (Ho et al., 1998; Rosenheck et al., 2006). An influential reconceptualization of the dopamine hypothesis in the 1990s proposed that sub-cortical hyperdopaminergia was secondary to cortical hypodopaminergia, in particular in frontal regions (Davis et al., 1991). The association between frontal dysfunction and cognitive impairment seen in schizophrenia is well established (Barch and Ceaser, 2012). Furthermore research in patients with chronic schizophrenia has shown that cortical dysfunction is linked to increased striatal dopamine dysfunction (Bertolino et al., 2000; Meyer-Lindenberg et al., 2002), in line with this hypothesis. However, as these studies were in chronic patients, it was not clear whether this reflected the long-term sequelae of the condition or was linked to the development of the disorder. More recent research in individuals showing prodromal indicators of schizophrenia has found that increased striatal dopamine synthesis capacity is associated with worse performance on cognitive tasks (Howes et al., 2009b), and that it is correlated with altered cortical function during cognitive tasks prior to the onset of schizophrenia as well (Fusar-poli et al., 2010; Allen et al., 2012a; Allen et al., 2012b). This indicates that a common factor could underlie both processes, but does not indicate which is primary, or what the common factor is. Whilst it remains plausible that the common factor is cortical dysfunction, animal models using mice that selectively overexpress striatal D2 receptors suggest that increased dopaminergic activity in the striatum may be able to account for both deficits in incentive motivation (Ward et al., 2012) and in cognitive functioning (Simpson et al., 2010), indicating that striatal dopaminergic dysfunction could contribute to the negative symptoms and cognitive impairments. These findings thus suggest that sub-cortical dopamine dysfunction could contribute to the cognitive and negative symptoms of schizophrenia as well as psychosis.

Limitations of the dopamine evidence in schizophrenia

The findings reviewed above provide evidence that there is a dopaminergic abnormality underlying schizophrenia, localise it to presynaptic dopamine dysfunction and link this to the symptoms of the disorder. However there are a number of findings that remain inadequately accounted for. These are discussed below.

Treatment resistance and non-dopaminergic forms of schizophrenia

One third of individuals with schizophrenia do not respond to non-clozapine antipsychotics (Mortimer et al., 2010) despite high levels of D2 occupancy (Kapur et al., 2000). Furthermore they do not respond to manipulations that deplete presynaptic dopamine either (Remington et al., 2012). The implication is that for a significant number of patients the pathophysiological basis of their symptoms involves more than dopaminergic excess, or may be unrelated to dopaminergic dysfunction. Demjaha et al. have shown that dopamine synthesis capacity appears was raised in individuals with a treatment responsive illness but not in treatment resistant patients (Demjaha et al., 2012). This is in agreement with earlier findings that found increased synaptic dopamine was predictive of treatment response (Abi-dargham et al., 2000). Overall this suggests that there may be a ‘non-dopaminergic’ sub-type of schizophrenia (Howes and Kapur, 2014).

Negative and cognitive symptoms

Whilst dopaminergic dysfunction has been linked to negative and cognitive symptoms, and there are plausible mechanisms to explain this (as reviewed above), the direction of causality has yet to be established (Howes et al., 2012a). Furthermore, the challenge studies provide evidence that dopamine elevation, albeit to supra-physiological levels, reduces negative symptoms (Laruelle et al., 1999), which is not consistent with a simple model of presynaptic dopamine dysregulation underlying negative symptoms. This could be accounted for by the hypothesis that there are regionally selective changes, with low cortical dopamine accounting for negative and cognitive symptoms (Laruelle, 2014). However, this hypothesis remains to be tested with in vivo imaging studies, although this is now possible (Narendran et al., 2009; Narendran et al., 2014). Notwithstanding this, in clinical practice the effects of dopamine antagonists and partial agonists on cognitive impairments and negative symptoms are modest at best (Murphy et al., 2006), and may even worsen cognitive function(Kim et al., 2013b). This implies that either the dopamine modulating tools currently employed are blunt instruments or that other pathways, such as those involving glutamate, contribute to cognitive dysfunction.

Dual diagnosis

Substance dependence is common in people with schizophrenia (Buckley et al., 2009) and a number of studies have examined this population. Decreased amphetamine induced dopamine release has been demonstrated in dual diagnosis patients compared to healthy controls (Thompson et al., 2013). Similarly, in clinically high risk individuals those that were cannabis dependent showed reduced dopamine release (Mizrahi et al., 2013). Imaging of dopamine synthesis capacity in individuals who experience psychotic symptoms when they smoke cannabis showed significantly reduced striatal [18F]-DOPA uptake compared to healthy volunteers (Bloomfield et al., 2014). These findings indicate that the link between substance abuse and psychosis may involve a pathway that is distinct from the striatal presynaptic dopamine dysfunction described above.

Glutamate

The NMDA receptor hypofunction hypothesis

Excitatory neurotransmission in the brain is primarily glutamatergic, with glutamatergic neurons utilising between 60 and 80 percent of total brain metabolic activity (Rothman et al., 2003). Glutamatergic neurotransmission occurs through metabotropic and ionotropic glutamate receptors, which are each subdivided into 3 groups. Group I metabotropic glutamate receptors (mGluR1 and mGluR5) are mainly postsynaptic, whereas group II (mGluR2 and mGluR3) and III (mGluR4, mGluR6, mGluR7 and mGluR8) are primarily presynaptic and modulate neurotransmitter release (Kew and Kemp, 2005). Ionotropic glutamate receptors are named after the agonists originally found to selectively activate them: a-amino-3-hydroxy-5-methyl-4-isoazolepropionic acid (AMPA), kainate and N-methyl- D-aspartate (NMDA) (Dingledine et al., 1999; Kew and Kemp, 2005).

The involvement of glutamatergic mechanisms in schizophrenia has been hypothesised for many years. There are reports as far back as 1949 of patients with schizophrenia being treated with glutamic acid (Kintzinger and Arnold, 1949). One of the earliest reports of glutamatergic abnormalities in living patients was the finding of reduced CSF glutamate levels in patients with schizophrenia (Kim et al., 1980). Despite early optimism, other groups failed to replicate this finding (Perry, 1982; Korpi et al., 1987). However, since that time, there has a growing body of evidence for alterations in other aspects of glutamatergic neurotransmission in schizophrenia. The glutamate hypothesis of schizophrenia originally formulated that there was a simple deficit in glutamatergic neurotransmission in the condition. This theory has been modified and developed over the years and, although many glutamate receptors have been implicated, the prevailing hypothesis is for the primary involvement of NMDA receptor dysfunction (Stone et al., 2007).

Indirect evidence for glutamate’s involvement

Post mortem and genetic studies

Post mortem studies provide evidence for alterations to glutamatergic functioning in schizophrenia. Studies of NMDA receptor expression in post mortem samples have produced some positive findings such as reduced NMDAR1 subunit density in the superior frontal cortex (Sokolov, 1998) and the superior temporal cortex (Humphries et al., 1996). However, overall findings regarding NMDA receptor density have been inconsistent (McCullumsmith et al., 2012). It seems that the abnormality in schizophrenia may primarily be aberrant glutamate receptor localisation as opposed to a generalised deficit (Hammond et al., 2014). This abnormality could arise as a result of changes in glutamate receptor trafficking molecules (Funk et al., 2009). Furthermore there is evidence for a variety of functional changes affecting the intracellular effects of the NMDA receptor that would have a major impact on glutamatergic signalling(Pitcher et al., 2011; Funk et al., 2012).

The potential role of glutamate in the pathophysiology of schizophrenia is also supported by recent genetics’ findings (Ripke et al., 2014). GRIN2A which codes for an NMDA receptor subunit was found to be associated with schizophrenia, as was SRR which plays a key role in pathways leading to the activation of the NMDA receptor.

NMDA receptor antagonists

The NMDA receptor dysfunction hypothesis of schizophrenia arose initially from the observation that non-competitive NMDA receptor antagonists, including phencyclidine (PCP), dizocilpine (MK-801) and ketamine, lead to immediate psychological effects, which closely resemble symptoms that occur in schizophrenia, including both positive and negative symptom domains (Krystal et al., 1994; Morgan and Curran, 2006; Javitt, 2007). Psychotic-like symptoms are also seen in chronic ketamine users (Morgan et al., 2009; Stone et al., 2014). These findings led to the use of NMDA receptor antagonists as a model system for schizophrenia. Olney and Farber pursued this approach in animal models, and put forward the first NMDA receptor dysfunction hypothesis of schizophrenia. They showed that animals given NMDA receptor antagonists developed neurotoxic changes in cortical brain regions, which they suggested were similar to reductions in brain volume seen in patients with schizophrenia (Olney and Farber, 1995). They found that AMPA antagonists (among other compounds) could block the downstream effects of NMDA receptor antagonism on neurotoxicity, and hypothesized that glutamate release might underlie the neurotoxic effects. This was subsequently confirmed in microdialysis studies (Moghaddam et al., 1997) and more recently with proton magnetic resonance spectroscopy (1H-MRS) studies following systemic ketamine administration in animals (Kim et al., 2011).

Further research demonstrated that injection of the NMDA receptor antagonist MK-801 directly into cortical regions did not lead to any evidence of neurodegenerative changes, whereas injection into anterior thalamus led to the same cortical changes as seen with systemic administration (Sharp et al., 2001), suggesting that the thalamus may be a critical site for NMDA receptor blockade in the generation of downstream effects by NMDA receptor antagonists, and, by extension, may also be a site of NMDA receptor dysfunction in schizophrenia (Olney and Farber, 1995; Stone et al., 2007).

The method of administration of NMDA antagonists may have relevance regarding which aspect of the schizophrenia syndrome one is attempting to model. Acute administration of NMDA antagonists can perhaps be viewed as a model of acute psychosis. In contrast, longer-term administration shows potential as a model for chronic schizophrenia. Sub chronic administration of PCP to rats has been shown to cause long lasting deficits in set shifting ability (Egerton et al., 2008), and reductions in functional connectivity (Dawson et al., 2014) similar to those observed in chronic schizophrenia. While longer term administration of ketamine to rats produced cognitive deficits similar to those observed in chronic schizophrenia (Featherstone et al., 2012). Furthermore, administration of NMDA antagonists early in life results in long-term behavioural and neuroanatomical effects (Harris et al., 2003) – providing a potential method for examining neurodevelopmental hypotheses.

The complete mechanism by which NMDA receptor antagonists lead to downstream increases in glutamate and neurotoxic changes in animals has not been fully elucidated, but evidence indicates it involves GABAergic interneurons. Specifically NMDA receptor antagonists have been shown to reduce GABAergic interneuron function (Homayoun and Moghaddam, 2007), and it is thought that this leads to an increase in pyramidal cell firing due to disinhibition (Olney and Farber, 1995). How NMDAR antagonism leads to reductions in GABAergic interneuron activity is less clear. It has been suggested that they may have preferential effects on NMDA receptors expressed on GABA neurons (Homayoun and Moghaddam, 2007), but this hypothesis has been contested (Rotaru et al., 2012). An alternative hypothesis is that NMDA receptor antagonist-induced changes in reactive oxygen species levels may be a central component of this mechanism since reducing superoxide levels prevented ketamine-induced changes in interneuron activity (Behrens et al., 2007). Furthermore, inhibiting reactive oxygen species formation blocks the behavioural effect of NMDA receptor antagonists in animals (Levkovitz et al., 2007; Zhang et al., 2007; Sorce et al., 2010; Monte et al., 2013).

Glutamatergic drugs treatments for schizophrenia

Findings from small clinical trials of glutamatergic drugs give further support to the NMDA receptor hypofunction hypothesis of schizophrenia (Stone, 2011). A number of studies have investigated NMDA receptor modulation, either via direct agonism at the glycineB modulatory site (using glycine or D-serine), or by increasing synaptic glycine levels through inhibition of glycine transporters (eg: sarcosine). A meta-analysis of published studies in 2009 reported that these drugs led to a significant improvement in residual positive and negative symptoms in patients with schizophrenia when administered in addition to existing antipsychotic treatment (Tsai and Lin, 2010). However, a subsequent, large clinical trial of D-serine, failed to show any benefit (Weiser et al., 2012).

Bitopertin, a glycine transporter inhibiter developed by Roche, showed promising efficacy for negative symptoms in an initial trial (Umbricht et al., 2014), and showed similar efficacy to olanzapine when used as monotherapy, in terms of patient readiness for discharge at 4 weeks compared to placebo (Bugarski-Kirola et al., 2014). However, two recent Phase III studies, showed no benefit of bitopertin for negative symptoms (Goff, 2014). Thus, whilst the results of an ongoing study of bitopertin for reducing positive symptoms in patients with schizophrenia are awaited, the results to date suggest the effect size is modest at most. Overall there is some evidence that modulation of the glycineB site may be therapeutic in schizophrenia but this this has not been fully established.

Minocycline, a tetracycline antibiotic, has more recently raised interest in light of its neuroprotective properties. An initial open label study (Miyaoka and Yasukawa, 2008) showed large improvements in both positive and negative symptoms. More rigorous controlled studies (Levkovitz et al., 2010; Chaudhry et al., 2012; Liu et al., 2014) examining its effect on negative symptoms have also shown clinically significant benefits. There are a number of potential mechanisms of action for minocycline, some of which involve the glutamatergic system. In animal models minocycline has been shown to counter the effects of multiple NMDA antagonists (Levkovitz et al., 2007; Zhang et al., 2007; Fujita et al., 2008). This effect on the glutamatergic system may be indirect and one possibility is that minocycline may modulate NMDA receptor signalling by inhibiting the formation of reactive oxygen species (Monte et al., 2013).

The striking results recently observed in patients with schizophrenia following a single dose of sodium nitroprusside (Hallak et al., 2013), may also have a glutamatergic mechanism underlying the clinical response. Preclinical work has shown that sodium nitroprusside is able to abolish the behavioural effects of phencyclidine (Bobanovic et al., 2000). The exact mechanism responsible for its clinical effects is not yet determined although there is some evidence that it may modulate NMDA receptor activity(Manzoni et al., 1992).

Another line of evidence comes from studies of drugs targeting downstream glutamate release, such as lamotrigine, LY2140023 and topirimate. Lamotrigine, a drug which reduces glutamate release, has been reported to inhibit ketamine-induced psychosis-like effects in healthy volunteers (Hosák and Libiger, 2002), and related changes in brain function measured using fMRI (Deakin et al., 2008). Early clinical data suggesting a benefit in partial responders to clozapine treatment (Dursun et al., 1999) has received subsequent support in a meta-analysis, although effects were relatively modest (Tiihonen et al., 2009).

LY2140023 is a drug developed by Eli Lilly as an agonist for presynaptic mGlu2/3 receptors to reduce glutamate release. It was found in an initial study to lead to significant improvements in positive and negative symptoms in patients with chronic schizophrenia (Patil et al., 2007). However, a subsequent phase II trial did not show any significant benefit over placebo, possibly due to a particularly high level of placebo response (Kinon et al., 2010).

Lastly, topiramate is a drug that may alter the downstream effect of excess glutamate through AMPA receptor antagonism. An initial positive open trial (Dursun and Deakin, 2001) using topirimate as an augmentation agent in treatment resistant schizophrenia was later replicated in a randomised controlled trial (Tiihonen et al., 2005). It has also been shown to reduce the behavioural effects of MK-801 in rats (Deutsch et al., 2002). However, as AMPA antagonism only occurs at higher doses, these inhibitory effects may occur through its enhancement of GABA transmission (Gibbs et al., 2000), potentially calling into question a direct glutamatergic mechanism for the drug.

Overall, whilst there is clearly some evidence that glutamatergic drugs are effective treatments for schizophrenia, findings are somewhat mixed and where supported by meta-analysis the effect size is modest. These studies thus provide some evidence to support the involvement of glutamatergic dysfunction in schizophrenia, but there is nothing like the weight of evidence for treatment effects that there is for dopamine receptor blockers.

Glutamate – In vivo imaging studies

Neuroimaging studies of glutamatergic function in psychosis began later than the studies on the dopamine system. Nevertheless there is a substantial body of evidence from studies of the effects of ketamine on brain function in healthy volunteers, and studies of brain glutamate levels in patients with prodromal and first episode psychosis, and with schizophrenia. These studies are summarised below, focusing on first episode and high risk studies as these ameliorate the risk of confounding by treatment effects (Poels et al., 2014).

SPECT Studies

There is only one published SPECT study of NMDA receptor binding in patients with schizophrenia to date (Pilowsky et al., 2006). This showed evidence of a relative deficit in NMDA receptor activity in the left hippocampus in unmedicated patients with schizophrenia (Pilowsky et al., 2006). The study is relatively small, and no other groups have attempted to replicate the findings, possibly because of concern around levels of non-specific binding associated with the tracer used (N-(1-napthyl)-N'-(3-[(123)I]-iodophenyl)-N-methylguanidine (123I-CNS 1261)), a problem that affects all NMDA receptor tracers currently available (Knol et al., 2009). Nevertheless it provides in vivo support for the NMDA hypofunction hypothesis, although this clearly requires further confirmation.

Proton Magnetic Resonance Spectroscopy Studies

Proton Magnetic Resonance Spectroscopy (1H-MRS) has been used to measure glutamate and glutamine levels in individuals at high risk of psychosis, as well as patients with first episode psychosis and chronic schizophrenia (Poels et al., 2014). Glutamine has been suggested to be a marker of glutamatergic neurotransmission, as it is generated after synaptic glutamate is taken up into astrocytes(Bak et al., 2006). 1H-MRS studies in schizophreniform psychosis and schizophrenia to date have generally found that individuals with clinical or familial risk, and those with first episode psychosis have increased glutamine in anterior cingulate cortex (Bartha et al., 1997; Théberge et al., 2002; Tibbo et al., 2004; Stone et al., 2009; Marsman et al., 2013; Tandon et al., 2013). The 1H-MRS studies of at risk or first episode populations are summarised in tables 4 and 5. In contrast, studies of patients with chronic schizophrenia have generally found normal or reduced cortical glutamate and glutamine levels (Block et al., 2000; Kegeles et al., 2000b; Théberge et al., 2003; Ohrmann et al., 2005; Ongur et al., 2009; Rowland et al., 2009; Tayoshi et al., 2009; Lutkenhoff et al., 2010; Natsubori et al., 2013). However, a recent large, well designed study found, that patients with chronic schizophrenia had elevated glutamine:glutamate ratio (Gln/Glu) in anterior cingulate cortex, with a correlation between frontal Gln/Glu and positive psychotic symptoms (Bustillo et al., 2014). Furthermore, in contrast to the data from a recently published meta-analysis (Marsman et al., 2013), Gln/Glu increased with the age of patients rather than decreased (Bustillo et al., 2014). The reason for this discrepancy is not clear, but may be due to differences in acquisition methodology, with this study employing a slightly longer echo time (40ms) for 1H-MRS acquisition (Bustillo et al., 2014).

Table 4.

The 1H-MRS studies in first episode schizophrenia

Study Medication status MN/MF/M/HC Phenotype Field strength Findings: Glu/Gln/Glx
MPFC Hipp Tha BG
Aoyama et al. 2011a 17/0/0/17 FEP 4.0T -/-/na -/↑/na
Bartha et al. 1997 10/0/0/10 FEP 1.5T -/↑/na
Bartha et al. 1999 11/0/0/11 FEP 1.5T -/-/na
Bustillo et al. 2010 14/0/0/10 FEP 4.0T -/-/na -/-/na
de la Fuente Sandoval et al. 2011b 36/0/0/40 FEP/UHR 3.0T ↑/na/-
de la Fuente Sandoval et al. 2013 24/0/0/18 FEP 3.0T ↑/na/↑
Galinska et al. 2009 1/0/29/19 FEP 1.5T na/na/- na/na/-
Gotoet al. 2012 c 16/18 FEP 3.0T na/na/↑
Olbrich et al. 2008 0/0/7/16 FEP 2.0T -/-/na
Théberge et al. 2002 21/0/0/21 FEP 4.0T -/↑/na -/↑/na

1H-MRS= Proton Magnetic Resonance Spectroscopy; MN- antipsychotic naïve; M – currently taking antipsychotic medication; HC – healthy control;FEP- first episode psychosis

Tha – Thalamus; BG – Basal Ganglia; Hipp – Hippocampus; MPFC – medial prefrontal cortex (including anterior cingulated cortex; Glx – Glutamate+ glutamine; Glu – glutamate; Gln – glutamine

↑ = significantly higher in patient group;v ↓ = significantly lower level in patient group; - = no significant difference; na – not analysed; UHR – Ultra high risk for psychosis

a

12 patients previously reported in Théberge et al. 2002

b

Subjects consist of 18 FEP and 18 UHR

c

The number of medicated vs. unmedicated patients is not clearly described.

Table 5.

The 1H-MRS studies in studies in individuals at increased risk of schizophrenia

Study Medication status MN/MF/M/HC Phenotype Field Strength Findings: Glu/Gln/Glx
MPFC Hipp Tha BG
Keshavan et al. 2009 40/0/0/46 UHR 1.5T na/na/- na/na/- na/na/-
Tandon et al. 2008 15/0/0/14 UHR 3.0T -/na/-
Stone et al. 2009 19/3/2/27 UHR 3.0T -/↑/- -/-/- ↓/-/↓
Yoo et al. 2009 22/0/0/22 UHR 1.5T na/na/- na/na/-
Lutkenhoff et al. 2010 12/0/0/21 Twins of Scz patients 3.0T ↓/na/na -/na/na
Tandon et al. 2013 23/0/0/24 Children of Scz parents 1.5T na/na/- na/na/↑ na/na/↑
Tibbo et al. 2004a 20/0/0/22 Children of Scz parents 3.0T na/na/na

1H-MRS= Proton Magnetic Resonance Spectroscopy; MN- antipsychotic naïve; M – currently taking antipsychotic medication; HC – healthy control; Scz - Schizophrenia

Tha – Thalamus; BG – Basal Ganglia; Hipp – Hippocampus; MPFC – medial prefrontal cortex (including anterior cingulated cortex; Glx – Glutamate+ glutamine; Glu – glutamate; Gln – glutamine

↑ = significantly higher in patient group;v ↓ = significantly lower level in patient group; - = no significant difference; na – not analysed; UHR – Ultra high risk for psychosis

a

Absolute values not reported but significantly increased glu:gln ration in MPFC of high risk group.

Another brain region with growing evidence for glutamatergic abnormalities in schizophrenia is the caudate nucleus. Glutamate and glutamine levels have been reported to be increased in individuals at familial risk for schizophrenia (Tandon et al., 2013), and individuals at clinical risk and unmedicated patients with first episode psychosis have been reported to have increased glutamate in this brain region (de la Fuente-Sandoval et al., 2011), although it appears to be affected by treatment, with four weeks of antipsychotic medication leading to a normalisation of caudate glutamate levels (de la Fuente-Sandoval et al., 2013).

There has been considerable interest in the possible utility of 1H-MRS in predicting treatment response in schizophrenia – with the theory that poor responders to conventional dopaminergic antipsychotic treatments may have more of a glutamatergic basis to their illness (Stone et al., 2010). There have now been a number of studies that give preliminary support to this hypothesis. We recently found that patients with first episode psychosis who were poor responders to antipsychotic drugs had increased levels of glutamate scaled to creatine (Glu/Cr) in anterior cingulate cortex compared to those patients who had responded (Egerton et al., 2012). Similarly we found that patients with established treatment resistant schizophrenia had increased anterior cingulate glutamate levels compared to controls and good responders (Demjaha et al., 2014).

One problem with these studies is that the data were acquired cross-sectionally, and so it is not possible to determine whether the elevated glutamate measures in anterior cingulate were associated with treatment resistance, or were simply a surrogate marker of increased psychopathology at the time of scanning. For example, it has been shown that panic induced by cholecystokinin tetrapeptide leads to an acute increase in measured glutamate in anterior cingulate (Zwanzger et al., 2013), and it is conceivable that a similar mechanism may occur in patients with symptoms of schizophrenia. However, a recent longitudinal study demonstrated that baseline Glx/Cr (reflecting the combination of GABA, glutamate and glutamine scaled to creatine) levels in the anterior cingulate of unmedicated patients predicted subsequent response to antipsychotic medication, with immediate responders having lower Glx/Cr (Szulc et al., 2013). Thus, although mental state may modulate glutamate levels in anterior cingulate, this measure may, nonetheless, be a useful marker in predicting antipsychotic response. Longitudinal studies in first episode patients are required to clarify this point.

Limitations of the glutamate evidence in schizophrenia

Although several lines of evidence point to there being glutamatergic abnormalities in schizophrenia, there are a number of potential limitations to the theory. The use of 1H-MRS as the primary tool for the in vivo imaging of the glutamatergic system has some limitations. In particular 1H-MRS may not be able to distinguish between intra and extracellular compartments – so changes could reflect alterations in either compartment. The development of specific radiotracers for imaging of NMDA and AMPA receptors(Miyake et al., 2011; Majo et al., 2013; McGinnity et al., 2014) will potentially enable the extracellular release of glutamate to be studied. 13C MRS is another potentially useful tool for allowing clearer understanding of the glutamatergic system(Mason et al., 2007). A further limitation of the glutamate hypothesis is that it is not clear exactly what NMDA hypofunction means at the molecular level. This partly reflects the limitations of the evidence available at the moment.

Another major shortcoming at present is that there are no glutamatergic drugs currently on the market for schizophrenia, and trials of glutamatergic treatments have not shown a conclusive or strong effect in most cases. Secondly, due to the close interactions between glutamate and dopamine, and the fact that ketamine has been reported to have effects on dopamine release, and may directly interact with high affinity D2 receptors(Kapur and Seeman, 2002; Egerton et al., 2012), it may not be possible to interpret some of the ketamine findings as purely glutamatergic. Thirdly, it is clear that schizophrenia is unlikely to arise from a single cause in all cases (Horvath and Mirnics, 2014; Howes and Kapur, 2014), and thus glutamate abnormalities may be present only in a subset of patients with the illness, and possibly only at a particular phase of illness in these individuals(Marsman et al., 2013). Finally, there is as yet no clear model to account for how the glutamate abnormalities observed could underlie the phenomenology of psychosis as seen in patients.

Integrating the dopamine and glutamate hypotheses

Whilst the evidence for the involvement of presynaptic dopamine dysfunction in the majority of cases of schizophrenia is compelling, dopamine dysfunction is most clearly linked to psychotic symptoms and the evidence for dopamine’s involvement in the negative and cognitive symptoms is much less clear-cut (Javitt and Zukin, 1991; Tamminga et al., 1995). In this respect glutamate models involving NMDA receptor blockade appear to be better able account for the range and nature of these aspects of schizophrenia (Javitt, 2010). A combination of both NMDA hypofunction and presynaptic dopamine dysfunction may therefore provide the best explanation of all the clinical aspects of schizophrenia.

The possibility that mesolimbic dopaminergic hyperactivity is secondary to diminished inhibitory control had been suggested before the glutamatergic system was explicitly identified as a candidate (Weinberger, 1987). Dopamine neurons are regulated by glutamatergic projections to the midbrain dopamine nuclei, which makes them potentially sensitive to changes in glutamatergic function (Miller and Abercrombie, 1996). This suggests that the dopamine function seen in schizophrenia could be secondary to altered glutamatergic function (McGuire et al., 2008) (Figure 1.). Supporting this, preclinical studies show that NMDA blockers such as ketamine and PCP change dopamine neuron firing patterns, and increase dopamine release to a challenge such as amphetamine (Miller and Abercrombie, 1996; Tsukada, 2000; Balla et al., 2003; Jackson et al., 2004). There is also in vivo imaging evidence that ketamine administration leads to increased dopamine release in humans, as indexed by change in D2/D3 receptor PET tracer binding (Breier et al., 1998; Smith et al., 1998; Vollenweider et al., 2000). Moreover baseline D2/D3 receptor availability is associated with increased sensitivity to the psychotogenic effects of ketamine in humans (Vernaleken et al., 2013), further supporting the close interaction between these two systems. However, other groups have found evidence for elevated dopamine release following ketamine only in individuals treated with amphetamine (Kegeles et al., 2000a; Aalto et al., 2002; Kegeles et al., 2002; Aalto et al., 2005), suggesting that ketamine’s effects may be more marked when the dopamine system is challenged. Thus, in schizophrenia NMDA hypofunction may make the dopamine system more sensitive to the effects of psychological stress.

Figure 1.

Figure 1

Interactions between glutamatergic and dopaminergic pathways

Whilst the studies discussed above provide evidence that dopamine dysregulation in schizophrenia could be secondary to glutamatergic dysfunction, they do not identify the specific brain circuits or regions involved. The prefrontal cortex and hippocampus have both been suggested as potential sites as both regulate midbrain dopamine neurons via glutamatergic projections to the midbrain (Christie et al., 1985; Grace, 1991; Sesack and Pickel, 1992).

Hippocampus

Studies using a rodent developmental model of schizophrenia show that altered hippocampal activity is associated with increased dopamine neuron population activity, and that inactivating the hippocampus reverses the dopamine alterations (Lodge and Grace, 2006). Altered hippocampal activity is also seen in patients with schizophrenia (Heckers et al., 1998; Harrison, 2004; Malaspina et al., 2004) and prodromal psychosis, and has been linked to the subsequent development of schizophrenia (Schobel et al., 2013). Furthermore, we recently reported in individuals with prodromal symptoms of psychosis had a negative correlation between hippocampal glutamate and striatal [18F]DOPA uptake, and this was most marked in those that went on to develop schizophrenia (Stone et al., 2010).

Prefrontal cortex

Whilst there are data linking altered frontal function to striatal presynaptic dopamine dysfunction in patients and at risk subjects (Bertolino et al., 2000; Meyer-Lindenberg et al., 2002; Fusar-Poli et al., 2011), we are not aware of any studies investigating the relationship between frontal glutamate to striatal dopamine dysfunction in patients.

There are thus several converging lines of data to suggest that glutamate dysfunction could underlie the dopamine dysfunction seen in schizophrenia and its prodrome, supporting an integrated hypothesis. However, as the data in patients are limited to date (Stone et al., 2010), this requires further testing. Furthermore, it is important to note that the glutamate and dopamine systems show extensive and reciprocal interactions, and thus it may be difficult to determine which is primary, with different authorities on the subject suggesting that both glutamate and dopamine drive abnormalities in the other system (Olney and Farber, 1995; Harrison and Weinberger, 2005; Coyle, 2006; Stone et al., 2007). Finally, an integrated hypothesis cannot readily explain the findings in treatment resistant patients discussed above, where glutamate dysfunction alone or other pathways may be critical.

Conclusions and Future Directions

The first decade or so of the 21st century has seen the accumulation of evidence that has allowed major refinements of the dopamine hypothesis of schizophrenia. We can now say that the major dopaminergic abnormality is presynaptic, it is present at the onset of illness and is related to the onset of psychosis. These findings appear robust: they are seen across methods and settings and, given their consistency, it would take a large body of new, contradictory evidence to bring them into doubt. Furthermore, we have plausible models of how presynaptic dopamine dysfunction might lead to the symptoms we see in the clinic, providing an explanation at both a neurobiological and clinical level. Nevertheless there are important limitations to our current understanding, in particular of how the dopamine changes might account for negative and cognitive symptoms. The 21st century has also seen continued empirical support for the glutamate hypothesis, with several converging lines of evidence indicating that a glutamatergic abnormality could underlie schizophrenia. However, whilst there have been undoubted advances in understanding the nature of glutamate dysfunction in schizophrenia, at present there are some inconsistencies and this has not led to significant advances in treatment. This should not be surprising given the lessons from the imaging studies of the dopamine system in schizophrenia, in particular that it takes time (more than two decades in the case of dopamine), and the application of a variety of imaging techniques to develop a clear understanding of the nature of the dysfunction affecting a system. Given the limitations of current approaches to imaging glutamate using MRS, progress requires the development of new methods to image the glutamate system. Notwithstanding these caveats, there are two possible explanations for the involvement of both dopamine and glutamate in schizophrenia. One is that they underlie different sub-types of the disorder, in line with the recent findings in treatment resistance. The other is an integrated hypothesis which could explain positive symptoms in terms of presynaptic dopamine, and negative and cognitive symptoms in terms of glutamate. Of course these are not mutually exclusive. Both possibilities require further testing in patient studies and this is likely to be needed to inform the development of new treatment approaches.

Abbreviations

= significantly higher in patient group

= significantly lower level in patient group

-

= no significant difference

[11C]-+-PHNO-

[11C]-(+)-4-propyl-9-hydroxynaphthoxazine;

[11C]DOPA

= L-[β-11C]Dihydroxyphenylalanine;

[123I]IBZM-

[123I](S)-(-)-3-iodo-2-hydroxy-6-methoxy-N- [(1-ethyl-2-pyrrolidinyl)methyl]benzamide

[18F]-DOPA- 6-

[18F]fluoro-L-Dihydroxyphenylalanine

1H-MRS

= Proton Magnetic Resonance Spectroscopy

5-HT2A

= 5-hydoxytryptamine (also known as serotonin) receptor sub-type 2A

AMPT

alpha-methyl-para-tyrosine

BG

Basal Ganglia

Chronic

Chronic psychotic illness

D1, D2, D3

= Dopamine receptor sub-type 1, 2 and 3 respectively

DAT

= Dopamine transporter

EPSE

= Extrapyramidal side-effects

FEP

first episode psychosis

Gln

glutamine

Glu

glutamate

Glx

Glutamate+ glutamine

HC

healthy control;

Hipp

Hippocampus

L-DOPA

= L-Dihydroxyphenylalanine

M

currently taking antipsychotic medication

MF

antipsychotic free

MIST

Montreal imaging stress task

MN

antipsychotic naïve

MPFC

medial prefrontal cortex (including anterior cingulated cortex

na

not analysed

PET

= Positron Emission Tomography

sd

standard deviation

SPECT

= Single Photon Emission Tomography

Tha

Thalamus

UHR

Ultra high risk

Footnotes

Declaration of interest: Declaration of interest: ODH and JMS have been on the speaker bureau or have received investigator-initiated grant support from a number of companies with an interest in antipsychotic drugs (AZ, BMS, Eli Lilly, Jansenn, Leyden-Delta, Lundbeck, Roche, Servier). This study was funded by a Medical Research Council (UK) grant to Dr Howes (grant number: MC-A656-5QD30) and the National Institute of Health Research Biomedical Research Council grant to King’s College London

References

  1. Aalto S, Hirvonen J, Kajander J, et al. Ketamine does not decrease striatal dopamine D2 receptor binding in man. Psychopharmacology (Berl) 2002;164:401–406. doi: 10.1007/s00213-002-1236-6. [DOI] [PubMed] [Google Scholar]
  2. Aalto S, Ihalainen J, Hirvonen J, et al. Cortical glutamate-dopamine interaction and ketamine-induced psychotic symptoms in man. Psychopharmacology (Berl) 2005;182:375–383. doi: 10.1007/s00213-005-0092-6. [DOI] [PubMed] [Google Scholar]
  3. Abi-Dargham a, Gil R, Krystal J, et al. Increased striatal dopamine transmission in schizophrenia: confirmation in a second cohort. Am J Psychiatry. 1998;155:761–7. doi: 10.1176/ajp.155.6.761. [DOI] [PubMed] [Google Scholar]
  4. Abi-Dargham A, Kegeles LS, Zea-Ponce Y, et al. Striatal amphetamine-induced dopamine release in patients with schizotypal personality disorder studied with single photon emission computed tomography and [123I]iodobenzamide. Biol Psychiatry. 2004;55:1001–6. doi: 10.1016/j.biopsych.2004.01.018. [DOI] [PubMed] [Google Scholar]
  5. Abi-Dargham A, Mawlawi O, Lombardo I, et al. Prefrontal dopamine D1 receptors and working memory in schizophrenia. J Neurosci. 2002;22:3708–19. doi: 10.1523/JNEUROSCI.22-09-03708.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Abi-dargham A, Rodenhiser J, Printz D, et al. Increased baseline occupancy of D2 receptors by dopamine in schizophrenia. Proc Natl Acad Sci U S A. 2000;97:8104–8109. doi: 10.1073/pnas.97.14.8104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Abi-Dargham A, van de Giessen E, Slifstein M, et al. Baseline and amphetamine-stimulated dopamine activity are related in drug-naïve schizophrenic subjects. Biol Psychiatry. 2009;65:1091–3. doi: 10.1016/j.biopsych.2008.12.007. [DOI] [PubMed] [Google Scholar]
  8. Abi-Dargham A, Xu X, Thompson 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. doi: 10.1177/0269881111409265. [DOI] [PubMed] [Google Scholar]
  9. Allen P, Chaddock Ca, Howes OD, et al. Abnormal relationship between medial temporal lobe and subcortical dopamine function in people with an ultra high risk for psychosis. Schizophr Bull. 2012a;38:1040–9. doi: 10.1093/schbul/sbr017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Allen P, Luigjes J, Howes OD, et al. Transition to psychosis associated with prefrontal and subcortical dysfunction in ultra high-risk individuals. Schizophr Bull. 2012b;38:1268–76. doi: 10.1093/schbul/sbr194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Andrews G. Cost-effectiveness of current and optimal treatment for schizophrenia. Br J Psychiatry. 2003;183:427–435. doi: 10.1192/bjp.183.5.427. [DOI] [PubMed] [Google Scholar]
  12. Aoyama N, Théberge J, Drost DJ, et al. Grey matter and social functioning correlates of glutamatergic metabolite loss in schizophrenia. Br J Psychiatry. 2011;198:448–56. doi: 10.1192/bjp.bp.110.079608. [DOI] [PubMed] [Google Scholar]
  13. Arnold A, Freeman H. Reserpine in Hospitalized Psychotics: A Controlled Study on Chronically Disturbed Women. AMA Arch Neurol psychiatry. 1956;76:281–285. doi: 10.1001/archneurpsyc.1956.02330270053012. [DOI] [PubMed] [Google Scholar]
  14. Bak LK, Schousboe A, Waagepetersen HS. The glutamate/GABA-glutamine cycle: aspects of transport, neurotransmitter homeostasis and ammonia transfer. J Neurochem. 2006;98:641–53. doi: 10.1111/j.1471-4159.2006.03913.x. [DOI] [PubMed] [Google Scholar]
  15. Balla A, Sershen H, Serra M, et al. Subchronic continuous phencyclidine administration potentiates amphetamine-induced frontal cortex dopamine release. Neuropsychopharmacology. 2003;28:34–44. doi: 10.1038/sj.npp.1300019. [DOI] [PubMed] [Google Scholar]
  16. Barch DM, Ceaser A. Cognition in schizophrenia: core psychological and neural mechanisms. Trends Cogn Sci. 2012;16:27–34. doi: 10.1016/j.tics.2011.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Baron JC, Martinot JL, Cambon H, et al. Striatal dopamine receptor occupancy during and following withdrawal from neuroleptic treatment: correlative evaluation by positron emission tomography and plasma prolactin levels. Psychopharmacology (Berl) 1989;99:463–472. doi: 10.1007/BF00589893. [DOI] [PubMed] [Google Scholar]
  18. Bartha R, al-Semaan YM, Williamson PC, et al. A short echo proton magnetic resonance spectroscopy study of the left mesial-temporal lobe in first-onset schizophrenic patients. Biol Psychiatry. 1999;45:1403–11. doi: 10.1016/s0006-3223(99)00007-4. [DOI] [PubMed] [Google Scholar]
  19. Bartha R, Williamson PC, Drost DJ, et al. Measurement of glutamate and glutamine in the medial prefrontal cortex of never-treated schizophrenic patients and healthy controls by proton magnetic resonance spectroscopy. Arch Gen Psychiatry. 1997;54:959–965. doi: 10.1001/archpsyc.1997.01830220085012. [DOI] [PubMed] [Google Scholar]
  20. Behrens MM, Ali SS, Dao DN, et al. Ketamine-induced loss of phenotype of fast-spiking interneurons is mediated by NADPH-oxidase. Science. 2007;318:1645–1647. doi: 10.1126/science.1148045. [DOI] [PubMed] [Google Scholar]
  21. Bertolino A, Breier A, Callicott JH, et al. The relationship between dorsolateral prefrontal neuronal N-acetylaspartate and evoked release of striatal dopamine in schizophrenia. Neuropsychopharmacology. 2000;22:125–132. doi: 10.1016/S0893-133X(99)00096-2. [DOI] [PubMed] [Google Scholar]
  22. Block W, Bayer TA, Tepest R, et al. Decreased frontal lobe ratio of N-acetyl aspartate to choline in familial schizophrenia: a proton magnetic resonance spectroscopy study. Neurosci Lett. 2000;289:147–151. doi: 10.1016/s0304-3940(00)01264-7. [DOI] [PubMed] [Google Scholar]
  23. Bloemen OJN, de Koning MB, Gleich T, et al. Striatal dopamine D2/3 receptor binding following dopamine depletion in subjects at Ultra High Risk for psychosis. Eur Neuropsychopharmacol. 2013;23:126–32. doi: 10.1016/j.euroneuro.2012.04.015. [DOI] [PubMed] [Google Scholar]
  24. Bloomfield M, Morgan CJ, Egerton A, et al. Dopaminergic function in cannabis users and its relationship to cannabis-induced psychotic symptoms. Biol Psychiatry. 2014;75:470–8. doi: 10.1016/j.biopsych.2013.05.027. P. [DOI] [PubMed] [Google Scholar]
  25. Bobanovic M, Bird DC, Robertson HA, Dursun SM. Blockade of phencyclidine-induced effects by a nitric oxide donor. Br J Pharmacol. 2000:1005–1012. doi: 10.1038/sj.bjp.0703406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Breier A, Adler CM, Weisenfeld N, et al. Effects of NMDA antagonism on striatal dopamine release in healthy subjects: application of a novel PET approach. Synapse. 1998;29:142–147. doi: 10.1002/(SICI)1098-2396(199806)29:2<142::AID-SYN5>3.0.CO;2-7. [DOI] [PubMed] [Google Scholar]
  27. Bromberg-Martin ES, Matsumoto M, Hikosaka O. Dopamine in motivational control: rewarding, aversive, and alerting. Neuron. 2010;68:815–34. doi: 10.1016/j.neuron.2010.11.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Brücke T, Podreka I, Angelberger P, et al. Dopamine D2 receptor imaging with SPECT: studies in different neuropsychiatric disorders. J Cereb Blood Flow Metab. 1991;11:220–228. doi: 10.1038/jcbfm.1991.53. [DOI] [PubMed] [Google Scholar]
  29. Brücke T, Roth J, Podreka I, et al. Striatal dopamine D2-receptor blockade by typical and atypical neuroleptics. Lancet. 1992;339:497. doi: 10.1016/0140-6736(92)91108-k. [DOI] [PubMed] [Google Scholar]
  30. Buckley PF, Miller BJ, Lehrer DS, Castle DJ. Psychiatric comorbidities and schizophrenia. Schizophr Bull. 2009;35:383–402. doi: 10.1093/schbul/sbn135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Bugarski-Kirola D, Wang A, Abi-Saab D, Blättler T. A phase II/III trial of bitopertin monotherapy compared with placebo in patients with an acute exacerbation of schizophrenia - Results from the CandleLyte study. Eur Neuropsychopharmacol. 2014 doi: 10.1016/j.euroneuro.2014.03.007. [DOI] [PubMed] [Google Scholar]
  32. Bustillo JR, Chen H, Jones T, et al. Increased Glutamine in Patients Undergoing Long-term Treatment for Schizophrenia⨌: A Proton Magnetic Resonance Spectroscopy Study at 3 T. JAMA psychiatry. 2014;71:265–72. doi: 10.1001/jamapsychiatry.2013.3939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Bustillo JR, Rowland LM, Mullins P, et al. 1H-MRS at 4 tesla in minimally treated early schizophrenia. Mol Psychiatry. 2010;15:629–36. doi: 10.1038/mp.2009.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Campden-Main B, Wegielski Z. The Control of Deviant Behaviour in Chronically Disturbed Psychotic Patients by the Oral Administration of Reserpine. Ann N Y Acad Sci. 1955;61:117–122. doi: 10.1111/j.1749-6632.1955.tb42458.x. [DOI] [PubMed] [Google Scholar]
  35. Carlsson A, Lindqvist M, Magnusson T. 3,4-Dihydroxyphenylalanine and 5-hydroxytryptophan as reserpine antagonists. Nature. 1957;180:1200. doi: 10.1038/1801200a0. [DOI] [PubMed] [Google Scholar]
  36. Carlsson A, Roos B, Wålinder J, Skott A. Further Studies on the Mechanism of Antipsychotic Action: Potentiation by ~t-Methyltyrosine of Thioridazine Effects in Chronic Schizophrenics. J Neural Transm. 1973;132:125–132. doi: 10.1007/BF01244665. [DOI] [PubMed] [Google Scholar]
  37. Catafau AM, Searle GE, Bullich S, et al. Imaging cortical dopamine D1 receptors using [11C]NNC112 and ketanserin blockade of the 5-HT 2A receptors. J Cereb Blood Flow Metab. 2010;30:985–93. doi: 10.1038/jcbfm.2009.269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Chaudhry IB, Hallak J, Husain N, et al. Minocycline benefits negative symptoms in early schizophrenia: a randomised double-blind placebo-controlled clinical trial in patients on standard treatment. J Psychopharmacol. 2012;26:1185–93. doi: 10.1177/0269881112444941. [DOI] [PubMed] [Google Scholar]
  39. Chen KC, Yang YK, Howes O, et al. Striatal dopamine transporter availability in drug-naive patients with schizophrenia: a case-control SPECT study with [(99m)Tc]-TRODAT-1 and a meta-analysis. Schizophr Bull. 2013;39:378–86. doi: 10.1093/schbul/sbr163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Christie MJ, Bridge S, James LB, Beart PM. Excitotoxin lesions suggest an aspartatergic projection from rat medial prefrontal cortex to ventral tegmental area. Brain Res. 1985;333:169–172. doi: 10.1016/0006-8993(85)90140-4. [DOI] [PubMed] [Google Scholar]
  41. Ciliax B, Heilman C. The Dopamine Transporter: Immunochemical characterization and Localization in Brain. J Neurosci. 1995;15:1714–1723. doi: 10.1523/JNEUROSCI.15-03-01714.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Coyle J. Glutamate and schizophrenia: beyond the dopamine hypothesis. Cell Mol Neurobiol. 2006;26:365–84. doi: 10.1007/s10571-006-9062-8. [DOI] [PubMed] [Google Scholar]
  43. Creese I, Burt D, Snyder S. Dopamine receptor binding predicts clinical and pharmacological potencies of antischizophrenic drugs. Science (80- ) 1976;192:481–483. doi: 10.1126/science.3854. [DOI] [PubMed] [Google Scholar]
  44. Dao-Castellana MH, Paillère-Martinot ML, Hantraye P, et al. Presynaptic dopaminergic function in the striatum of schizophrenic patients. Schizophr Res. 1997;23:167–74. doi: 10.1016/S0920-9964(96)00102-8. [DOI] [PubMed] [Google Scholar]
  45. Davis KL, Kahn RS, Ko G, Davidson M. Dopamine in schizophrenia: a review and reconceptualization. Am J Psychiatry. 1991;148:1474–1486. doi: 10.1176/ajp.148.11.1474. [DOI] [PubMed] [Google Scholar]
  46. Dawson N, Xiao X, McDonald M, et al. Sustained NMDA receptor hypofunction induces compromised neural systems integration and schizophrenia-like alterations in functional brain networks. Cereb Cortex. 2014;24:452–64. doi: 10.1093/cercor/bhs322. [DOI] [PubMed] [Google Scholar]
  47. De la Fuente-Sandoval C, León-Ortiz P, Azcárraga M, et al. Glutamate levels in the associative striatum before and after 4 weeks of antipsychotic treatment in first-episode psychosis: a longitudinal proton magnetic resonance spectroscopy study. JAMA psychiatry. 2013;70:1057–66. doi: 10.1001/jamapsychiatry.2013.289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. De la Fuente-Sandoval C, León-Ortiz P, Favila R, et al. Higher levels of glutamate in the associative-striatum of subjects with prodromal symptoms of schizophrenia and patients with first-episode psychosis. Neuropsychopharmacology. 2011;36:1781–1791. doi: 10.1038/npp.2011.65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Deakin JFW, Lees J, McKie S, et al. Glutamate and the neural basis of the subjective effects of ketamine: a pharmaco-magnetic resonance imaging study. Arch Gen Psychiatry. 2008;65:154–164. doi: 10.1001/archgenpsychiatry.2007.37. [DOI] [PubMed] [Google Scholar]
  50. Demjaha A, Egerton A, Murray RM, et al. Antipsychotic treatment resistance in schizophrenia associated with elevated glutamate levels but normal dopamine function. Biol Psychiatry. 2014;75:e11–3. doi: 10.1016/j.biopsych.2013.06.011. [DOI] [PubMed] [Google Scholar]
  51. Demjaha A, Murray RM, McGuire PK, et al. Dopamine synthesis capacity in patients with treatment-resistant schizophrenia. Am J Psychiatry. 2012;169:1203–10. doi: 10.1176/appi.ajp.2012.12010144. [DOI] [PubMed] [Google Scholar]
  52. Deutsch SI, Rosse RB, Billingslea EN, et al. Topiramate antagonizes MK-801 in an animal model of schizophrenia. Eur J Pharmacol. 2002;449:121–125. doi: 10.1016/s0014-2999(02)02041-1. [DOI] [PubMed] [Google Scholar]
  53. Dingledine R, Borges K, Bowie D, Traynelis SF. The glutamate receptor ion channels. Pharmacol Rev. 1999;51:7–61. [PubMed] [Google Scholar]
  54. Dursun S, McIntosh D, Milliken H. Clozapine plus lamotrigine in treatment-resistant schizophrenia. Arch Gen. 1999;245:372–3. doi: 10.1001/archpsyc.56.10.950. [DOI] [PubMed] [Google Scholar]
  55. Dursun SM, Deakin JFW. Augmenting antipsychotic treatment with lamotrigine or topiramate in patients with treatment-resistant schizophrenia: a naturalistic caseseries outcome study. J Psychopharmacol. 2001;15:297–301. doi: 10.1177/026988110101500409. [DOI] [PubMed] [Google Scholar]
  56. Egerton A, Brugger S, Raffin M, et al. Anterior Cingulate Glutamate Levels Related to Clinical Status Following Treatment in First-Episode Schizophrenia. Neuropsychopharmacology. 2012;37:2515–2521. doi: 10.1038/npp.2012.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Egerton A, Chaddock Ca, Winton-Brown TT, et al. Presynaptic striatal dopamine dysfunction in people at ultra-high risk for psychosis: findings in a second cohort. Biol Psychiatry. 2013;74:106–12. doi: 10.1016/j.biopsych.2012.11.017. [DOI] [PubMed] [Google Scholar]
  58. Egerton A, Demjaha A, McGuire P, et al. The test-retest reliability of 18F-DOPA PET in assessing striatal and extrastriatal presynaptic dopaminergic function. Neuroimage. 2010a;50:524–531. doi: 10.1016/j.neuroimage.2009.12.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Egerton A, Hirani E, Ahmad R, et al. Further evaluation of the carbon11-labeled D(2/3) agonist PET radiotracer PHNO: reproducibility in tracer characteristics and characterization of extrastriatal binding. Synapse. 2010a;64:301–12. doi: 10.1002/syn.20718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Egerton A, Reid L, McGregor S, et al. Subchronic and chronic PCP treatment produces temporally distinct deficits in attentional set shifting and prepulse inhibition in rats. Psychopharmacology (Berl) 2008;198:37–49. doi: 10.1007/s00213-008-1071-5. [DOI] [PubMed] [Google Scholar]
  61. 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. doi: 10.1007/s11307-007-0077-4. [DOI] [PubMed] [Google Scholar]
  62. Elkashef A, Doudet D, Bryant T. 6- 18 F-DOPA PET study in patients with schizophrenia. Psychiatry Res Neuroimaging. 2000;100:1–11. doi: 10.1016/s0925-4927(00)00064-0. [DOI] [PubMed] [Google Scholar]
  63. Farde L, Wiesel FA, Jansson P, et al. An open label trial of raclopride in acute schizophrenia. Confirmation of D2-dopamine receptor occupancy by PET. Psychopharmacology (Berl) 1988;94:1–7. doi: 10.1007/BF00735871. [DOI] [PubMed] [Google Scholar]
  64. Featherstone RE, Liang Y, Saunders Ja, et al. Subchronic ketamine treatment leads to permanent changes in EEG cognition and the astrocytic glutamate transporter EAAT2 in mice. Neurobiol Dis. 2012;47:338–46. doi: 10.1016/j.nbd.2012.05.003. [DOI] [PubMed] [Google Scholar]
  65. Fujita Y, Ishima T, Kunitachi S, et al. Phencyclidine-induced cognitive deficits in mice are improved by subsequent subchronic administration of the antibiotic drug minocycline. Prog Neuropsychopharmacol Biol Psychiatry. 2008;32:336–9. doi: 10.1016/j.pnpbp.2007.08.031. [DOI] [PubMed] [Google Scholar]
  66. Funk A, Rumbaugh G, Harotunianc V, et al. Decreased expression of NMDA receptor-associated proteins in frontal cortex of elderly patients with schizophrenia. Neuroreport. 2009;20:1019–1022. doi: 10.1097/WNR.0b013e32832d30d9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Funk AJ, McCullumsmith RE, Haroutunian V, Meador-Woodruff JH. Abnormal activity of the MAPK- and cAMP-associated signaling pathways in frontal cortical areas in postmortem brain in schizophrenia. Neuropsychopharmacology. 2012;37:896–905. doi: 10.1038/npp.2011.267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Fusar-Poli P, Howes O, Valmaggia L, McGuire P. “Truman” signs and vulnerability to psychosis. Br J Psychiatry. 2008;193:168. doi: 10.1192/bjp.193.2.168. [DOI] [PubMed] [Google Scholar]
  69. Fusar-Poli P, Howes OD, Allen P, et al. Abnormal prefrontal activation directly related to pre-synaptic striatal dopamine dysfunction in people at clinical high risk for psychosis. Mol Psychiatry. 2011;16:67–75. doi: 10.1038/mp.2009.108. [DOI] [PubMed] [Google Scholar]
  70. Fusar-poli P, Howes OD, Allen P, et al. Abnormal Frontostriatal Interactions in People With Prodromal Signs of Psychosis. Arch Gen Psychiatry. 2010;67:683–691. doi: 10.1001/archgenpsychiatry.2010.77. [DOI] [PubMed] [Google Scholar]
  71. Galińska B, Szulc A, Tarasów E, et al. Duration of untreated psychosis and proton magnetic resonance spectroscopy (1H-MRS) findings in first-episode schizophrenia. Med Sci Monit. 2009;15:CR82–88. [PubMed] [Google Scholar]
  72. Garety Pa, Kuipers E, Fowler D, et al. A cognitive model of the positive symptoms of psychosis. Psychol Med. 2001;31:189–95. doi: 10.1017/s0033291701003312. [DOI] [PubMed] [Google Scholar]
  73. Gibbs JW, Sombati S, DeLorenzo RJ, Coulter DA. Cellular actions of topiramate: blockade of kainate-evoked inward currents in cultured hippocampal neurons. Epilepsia. 2000;41(Suppl 1):S10–S16. doi: 10.1111/j.1528-1157.2000.tb02164.x. [DOI] [PubMed] [Google Scholar]
  74. Goff D. Bitopertin: The Good News and Bad News. JAMA psychiatry. 2014 doi: 10.1001/jamapsychiatry.2014.163.2. [DOI] [PubMed] [Google Scholar]
  75. Goto N, Yoshimura R, Kakeda S, et al. Six-month treatment with atypical antipsychotic drugs decreased frontal-lobe levels of glutamate plus glutamine in early-stage first-episode schizophrenia. Neuropsychiatr Dis Treat. 2012;8:119–22. doi: 10.2147/NDT.S25582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Goyer PF, Berridge MS, Morris ED, et al. PET measurement of neuroreceptor occupancy by typical and atypical neuroleptics. J Nucl Med. 1996;37:1122–1127. [PubMed] [Google Scholar]
  77. Grace AA. Phasic Versus tonic Dopamine Release and the Modulation of Dopamine System Responsivity: A Hypohtesis for the Etiology of Schizophrenia. Neuroscience. 1991;41:1–24. doi: 10.1016/0306-4522(91)90196-u. [DOI] [PubMed] [Google Scholar]
  78. Graff-Guerrero A, Mizrahi R, Agid O, et al. The dopamine D2 receptors in high-affinity state and D3 receptors in schizophrenia: a clinical [11C]-(+)-PHNO PET stud. Neuropsychopharmacology. 2009;34:1078–86. doi: 10.1038/npp.2008.199. [DOI] [PubMed] [Google Scholar]
  79. Gustavsson A, Svensson M, Jacobi F, et al. Cost of disorders of the brain in Europe 2010. Eur Neuropsychopharmacol. 2011;21:718–79. doi: 10.1016/j.euroneuro.2011.08.008. [DOI] [PubMed] [Google Scholar]
  80. Hallak JEC, Maia-de-Oliveira JP, Abrao J, et al. Rapid improvement of acute schizophrenia symptoms after intravenous sodium nitroprusside: a randomized, double-blind, placebo-controlled trial. JAMA psychiatry. 2013;70:668–76. doi: 10.1001/jamapsychiatry.2013.1292. [DOI] [PubMed] [Google Scholar]
  81. Hammond J, Shan D, Meador-Woodruff J, McCullumsmith R. Evidence of Glutamatergic Dysfunction in the Pathophysiology of Schizophrenia. In: Popoli M, Diamond D, Sanacora G, editors. Synaptic Stress Pathog Neuropsychiatr Disord. Springer New York; New York, NY: 2014. pp. 265–294. [Google Scholar]
  82. Harris LW, Sharp T, Gartlon J, et al. Long-term behavioural, molecular and morphological effects of neonatal NMDA receptor antagonism. Eur J Neurosci. 2003;18:1706–1710. doi: 10.1046/j.1460-9568.2003.02902.x. [DOI] [PubMed] [Google Scholar]
  83. Harrison PJ. The hippocampus in schizophrenia: a review of the neuropathological evidence and its pathophysiological implications. Psychopharmacology (Berl) 2004;174:151–62. doi: 10.1007/s00213-003-1761-y. [DOI] [PubMed] [Google Scholar]
  84. Harrison PJ, Weinberger DR. Schizophrenia genes, gene expression, and neuropathology: on the matter of their convergence. Mol Psychiatry. 2005;10:40–68. doi: 10.1038/sj.mp.4001558. image 5. [DOI] [PubMed] [Google Scholar]
  85. Heckers S, Rauch SL, Goff D, et al. Impaired recruitment of the hippocampus during conscious recollection in schizophrenia. Nat Neurosci. 1998;1:318–23. doi: 10.1038/1137. [DOI] [PubMed] [Google Scholar]
  86. Heinz A, Schlagenhauf F. Dopaminergic dysfunction in schizophrenia: salience attribution revisited. Schizophr Bull. 2010;36:472–485. doi: 10.1093/schbul/sbq031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Hietala J, Syvälahti E, Vilkman H, et al. Depressive symptoms and presynaptic dopamine function in neuroleptic- naive schizophrenia. Schizophr Res. 1999;35:41–50. doi: 10.1016/s0920-9964(98)00113-3. [DOI] [PubMed] [Google Scholar]
  88. Hietala J, Syvälahti E, Vuorio K, et al. Presynaptic dopamine function in striatum of neuroleptic-naive schizophrenic patients. Lancet. 1995;346:1130–1131. doi: 10.1016/s0140-6736(95)91801-9. [DOI] [PubMed] [Google Scholar]
  89. 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. doi: 10.1176/ajp.2006.163.10.1747. [DOI] [PubMed] [Google Scholar]
  90. Ho BC, Nopoulos P, Flaum M, et al. Two-year outcome in first-episode schizophrenia: predictive value of symptoms for quality of life. Am J Psychiatry. 1998;155:1196–201. doi: 10.1176/ajp.155.9.1196. [DOI] [PubMed] [Google Scholar]
  91. Homayoun H, Moghaddam B. NMDA receptor hypofunction produces opposite effects on prefrontal cortex interneurons and pyramidal neurons. J Neurosci. 2007;27:11496–11500. doi: 10.1523/JNEUROSCI.2213-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Horvath S, Mirnics K. Schizophrenia as a Disorder of Molecular Pathways. Biol Psychiatry. 2014 doi: 10.1016/j.biopsych.2014.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Hosák L, Libiger J. Antiepileptic drugs in schizophrenia: a review. Eur Psychiatry. 2002;17:371–378. doi: 10.1016/s0924-9338(02)00696-x. [DOI] [PubMed] [Google Scholar]
  94. Howes O, Bose S, Turkheimer F, et al. Progressive increase in striatal dopamine synthesis capacity as patients develop psychosis⨌: a PET study. Mol Psychiatry. 2011a;16:885–886. doi: 10.1038/mp.2011.20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Howes O, Bose S, Turkheimer FE, et al. Dopamine synthesis capacity before onset of psychosis: a prospective -DOPA PET imaging study. Am J Psychiatry. 2011b;168:1311–1317. doi: 10.1176/appi.ajp.2011.11010160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Howes O, Egerton A, Allan V. Mechanisms underlying psychosis and antipsychotic treatment response in schizophrenia: insights from PET and SPECT imaging. Curr. 2009a;15:2550–2559. doi: 10.2174/138161209788957528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Howes O, Fusar-Poli P, Bloomfield M, et al. From the Prodrome to Chronic Schizophrenia: the neurobiology underlying psychotic symptoms and cognitive impairments. Curr Pharm Des. 2012a;18:459–465. doi: 10.2174/138161212799316217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Howes O, Murray R. Schizophrenia: an integrated sociodevelopmental-cognitive model. Lancet. 2014;6736:1–11. doi: 10.1016/S0140-6736(13)62036-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Howes O, Williams M, Ibrahim K, Leung G. Midbrain dopamine function in schizophrenia and depression: a post-mortem and positron emission tomographic imaging study. Brain. 2013a:3242–3251. doi: 10.1093/brain/awt264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Howes OD, Kambeitz J, Kim E, et al. The Nature of Dopamine Dysfunction in Schizophrenia and What This Means for Treatment. Arch Gen Psychiatry. 2012b;69:776–786. doi: 10.1001/archgenpsychiatry.2012.169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Howes OD, Kapur S. The dopamine hypothesis of schizophrenia: version III--the final common pathway. Schizophr Bull. 2009;35:549–62. doi: 10.1093/schbul/sbp006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Howes OD, Kapur S. Br J Psychiatry. 2014 doi: 10.1192/bjp.bp.113.138578. In Press. [DOI] [PubMed] [Google Scholar]
  103. Howes OD, Montgomery AJ, Asselin M-C, et al. Elevated striatal dopamine function linked to prodromal signs of schizophrenia. Arch Gen Psychiatry. 2009b;66:13–20. doi: 10.1001/archgenpsychiatry.2008.514. [DOI] [PubMed] [Google Scholar]
  104. Howes OD, Shotbolt P, Bloomfield M, et al. Dopaminergic function in the psychosis spectrum: an [18F]-DOPA imaging study in healthy individuals with auditory hallucinations. Schizophr Bull. 2013b;39:807–14. doi: 10.1093/schbul/sbr195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Humphries C, Mortimer A, Hirsch S, de Belleroche J. NMDA receptor mRNA correlation with antemortem cognitive impairment in schizophrenia. Neuroreport. 1996;7:2051–5. doi: 10.1097/00001756-199608120-00040. [DOI] [PubMed] [Google Scholar]
  106. Huttunen J, Heinimaa M, Svirskis T, et al. Striatal dopamine synthesis in first-degree relatives of patients with schizophrenia. Biol Psychiatry. 2008;63:114–7. doi: 10.1016/j.biopsych.2007.04.017. [DOI] [PubMed] [Google Scholar]
  107. Jablensky A. Epidemiology of schizophrenia: the global burden of disease and disability. Eur Arch Psychiatry Clin Neurosci. 2000;250:274–85. doi: 10.1007/s004060070002. [DOI] [PubMed] [Google Scholar]
  108. Jablensky A, Sartorius N, Ernberg G, et al. Schizophrenia: manifestations, incidence and course in different cultures. A World Health Organization ten-country study. Psychol Med Monogr. 1992;20(Suppl):1–97. doi: 10.1017/s0264180100000904. [DOI] [PubMed] [Google Scholar]
  109. Jackson ME, Homayoun H, Moghaddam B. NMDA receptor hypofunction produces concomitant firing rate potentiation and burst activity reduction in the prefrontal cortex. Proc Natl Acad Sci U S A. 2004;101:8467–72. doi: 10.1073/pnas.0308455101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Javitt D, Zukin S. Recent Advances in the Phencyclidine Model of Schizophrenia. Am J Psychiatry. 1991;148:1301–1308. doi: 10.1176/ajp.148.10.1301. [DOI] [PubMed] [Google Scholar]
  111. Javitt DC. Glutamatergic theories of schizophrenia. Isr J Psychiatry Relat Sci. 2010;47:4–16. [PubMed] [Google Scholar]
  112. Javitt DC. Glutamate and Schizophrenia: Phencyclidine, N-Methyl-d-Aspartate Receptors, and Dopamine-Glutamate Interactions. Int Rev Neurobiol. 2007;78:69–108. doi: 10.1016/S0074-7742(06)78003-5. [DOI] [PubMed] [Google Scholar]
  113. Juckel G, Schlagenhauf F, Koslowski M, et al. Dysfunction of ventral striatal reward prediction in schizophrenia. Neuroimage. 2006;29:409–16. doi: 10.1016/j.neuroimage.2005.07.051. [DOI] [PubMed] [Google Scholar]
  114. Kaalund SS, Newburn EN, Ye T, et al. Contrasting changes in DRD1 and DRD2 splice variant expression in schizophrenia and affective disorders, and associations with SNPs in postmortem brain. Mol Psychiatry. 2013 doi: 10.1038/mp.2013.165. [DOI] [PubMed] [Google Scholar]
  115. Kambeitz J, Abi-Dargham A, Kapur S, Howes O. Alterations in cortical and extrastriatal subcortical dopamine function in schizophrenia: systematic review and meta-analysis of imaging studies. Br J Psychiatry. 2014;204:420–429. doi: 10.1192/bjp.bp.113.132308. [DOI] [PubMed] [Google Scholar]
  116. Kapur S. Psychosis as a State of Aberrant Salience : A framework linking biology, phenomenology, and Pharmacology in Schizophrenia. Am J Psychiatry. 2003;160:13–23. doi: 10.1176/appi.ajp.160.1.13. [DOI] [PubMed] [Google Scholar]
  117. Kapur S, Seeman P. NMDA receptor antagonists ketamine and PCP have direct effects on the dopamine D(2) and serotonin 5-HT(2)receptors-implications for models of schizophrenia. Mol Psychiatry. 2002;7:837–844. doi: 10.1038/sj.mp.4001093. [DOI] [PubMed] [Google Scholar]
  118. Kapur S, Zipursky R, Jones C, et al. Relationship between dopamine D(2) occupancy, clinical response, and side effects: a double-blind PET study of first-episode schizophrenia. Am J Psychiatry. 2000;157:514–20. doi: 10.1176/appi.ajp.157.4.514. [DOI] [PubMed] [Google Scholar]
  119. Kapur S, Zipursky RB, Jones C, et al. The D2 receptor occupancy profile of loxapine determined using PET. Neuropsychopharmacology. 1996;15:562–566. doi: 10.1016/S0893-133X(96)00100-5. [DOI] [PubMed] [Google Scholar]
  120. Karlsson P, Farde L, Halldin C, Sedvall G. PET study of D(1) dopamine receptor binding in neuroleptic-naive patients with schizophrenia. Am J Psychiatry. 2002;159:761–7. doi: 10.1176/appi.ajp.159.5.761. [DOI] [PubMed] [Google Scholar]
  121. Kegeles LS, Abi-Dargham a, Zea-Ponce Y, et al. Modulation of amphetamine-induced striatal dopamine release by ketamine in humans: implications for schizophrenia. Biol Psychiatry. 2000a;48:627–40. doi: 10.1016/s0006-3223(00)00976-8. [DOI] [PubMed] [Google Scholar]
  122. Kegeles LS, Abi-Dargham A, Frankle WG, et al. Increased synaptic dopamine function in associative regions of the striatum in schizophrenia. Arch Gen Psychiatry. 2010;67:231–9. doi: 10.1001/archgenpsychiatry.2010.10. [DOI] [PubMed] [Google Scholar]
  123. Kegeles LS, Martinez D, Kochan LD, et al. NMDA antagonist effects on striatal dopamine release: positron emission tomography studies in humans. Synapse. 2002;43:19–29. doi: 10.1002/syn.10010. [DOI] [PubMed] [Google Scholar]
  124. Kegeles LS, Shungu DC, Anjilvel S, et al. Hippocampal pathology in schizophrenia: magnetic resonance imaging and spectroscopy studies. Psychiatry Res. 2000b;98:163–175. doi: 10.1016/s0925-4927(00)00044-5. [DOI] [PubMed] [Google Scholar]
  125. Kew JNC, Kemp JA. Ionotropic and metabotropic glutamate receptor structure and pharmacology. Psychopharmacology (Berl) 2005;179:4–29. doi: 10.1007/s00213-005-2200-z. [DOI] [PubMed] [Google Scholar]
  126. Kim E, Howes OD, Kapur S. Molecular imaging as a guide for the treatment of central nervous system disorders. Dialogues Clin Neurosci. 2013a;15:315–28. doi: 10.31887/DCNS.2013.15.3/ekim. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Kim E, Howes OD, Turkheimer FE, et al. The relationship between antipsychotic D2 occupancy and change in frontal metabolism and working memory : A dual [(11)C]raclopride and [(18) F]FDG imaging study with aripiprazole. Psychopharmacology (Berl) 2013b;227:221–9. doi: 10.1007/s00213-012-2953-0. [DOI] [PubMed] [Google Scholar]
  128. Kim JS, Kornhuber HH, Schmid-Burgk W, Holzmüller B. Low cerebrospinal fluid glutamate in schizophrenic patients and a new hypothesis on schizophrenia. Neurosci Lett. 1980;20:379–382. doi: 10.1016/0304-3940(80)90178-0. [DOI] [PubMed] [Google Scholar]
  129. Kim S, Lee H, Kim H-J, et al. In vivo and ex vivo evidence for ketamine-induced hyperglutamatergic activity in the cerebral cortex of the rat: Potential relevance to schizophrenia. NMR Biomed. 2011;24:1235–42. doi: 10.1002/nbm.1681. [DOI] [PubMed] [Google Scholar]
  130. Kinon BJ, Zhang L, Williams JE, et al. LY2140023 monohydrate: An agonist at the MLGU2/3 receptor for the treatment of schizophrenia. Schizophr Res. 2010;117:379–379. [Google Scholar]
  131. Kintzinger H, Arnold DG. A preliminary study of the effects of glutamic acid on catatonic schizophrenics. Rorschach Res Exch J Proj Tech. 1949;13:210–8. doi: 10.1080/10683402.1949.10381459. [DOI] [PubMed] [Google Scholar]
  132. Klemm E, Grünwald F, Kasper S, et al. [123I]IBZM SPECT for imaging of striatal D2 dopamine receptors in 56 schizophrenic patients taking various neuroleptics. Am J Psychiatry. 1996;153:183–190. doi: 10.1176/ajp.153.2.183. [DOI] [PubMed] [Google Scholar]
  133. Knol RJJ, de Bruin K, van Eck-Smit BLF, et al. In vivo [(123)I]CNS-1261 binding to D-serine-activated and MK801-blocked NMDA receptors: A storage phosphor imaging study in rats. Synapse. 2009;63:557–564. doi: 10.1002/syn.20629. [DOI] [PubMed] [Google Scholar]
  134. Korpi ER, Kaufmann CA, Marnela KM, Weinberger DR. Cerebrospinal fluid amino acid concentrations in chronic schizophrenia. Psychiatry Res. 1987;20:337–345. doi: 10.1016/0165-1781(87)90095-3. [DOI] [PubMed] [Google Scholar]
  135. 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. doi: 10.1016/j.lfs.2010.03.018. [DOI] [PubMed] [Google Scholar]
  136. Krystal JH, Karper LP, Seibyl JP, et al. Subanesthetic effects of the noncompetitive NMDA antagonist, ketamine, in humans. Psychotomimetic, perceptual, cognitive, and neuroendocrine responses. Arch Gen Psychiatry. 1994;51:199–214. doi: 10.1001/archpsyc.1994.03950030035004. [DOI] [PubMed] [Google Scholar]
  137. Kumakura Y, Cumming P. PET studies of cerebral levodopa metabolism: a review of clinical findings and modeling approaches. Neuroscientist. 2009;15:635–50. doi: 10.1177/1073858409338217. [DOI] [PubMed] [Google Scholar]
  138. Kumakura Y, Cumming P, Vernaleken I, et al. Elevated [18F]fluorodopamine turnover in brain of patients with schizophrenia: an [18F]fluorodopa/positron emission tomography study. J Neurosci. 2007;27:8080–7. doi: 10.1523/JNEUROSCI.0805-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Laruelle M. Schizophrenia: from dopaminergic to glutamatergic interventions. Curr Opin Pharmacol. 2014;14C:97–102. doi: 10.1016/j.coph.2014.01.001. [DOI] [PubMed] [Google Scholar]
  140. Laruelle M, Abi-Dargham a, Gil R, et al. Increased dopamine transmission in schizophrenia: relationship to illness phases. Biol Psychiatry. 1999;46:56–72. doi: 10.1016/s0006-3223(99)00067-0. [DOI] [PubMed] [Google Scholar]
  141. Levkovitz Y, Levi U, Braw Y, Cohen H. Minocycline, a second-generation tetracycline, as a neuroprotective agent in an animal model of schizophrenia. Brain Res. 2007;1154:154–162. doi: 10.1016/j.brainres.2007.03.080. [DOI] [PubMed] [Google Scholar]
  142. Levkovitz Y, Mendlovich S, Riwkes S, et al. A double-blind, randomized study of minocycline for the treatment of negative and cognitive symptoms in early-phase schizophrenia. J Clin Psychiatry. 2010;71:138–49. doi: 10.4088/JCP.08m04666yel. [DOI] [PubMed] [Google Scholar]
  143. Lieberman Ja, Kinon BJ, Loebel aD. Dopaminergic mechanisms in idiopathic and drug-induced psychoses. Schizophr Bull. 1990;16:97–110. doi: 10.1093/schbul/16.1.97. [DOI] [PubMed] [Google Scholar]
  144. Lieberman JA, Kane JM, Alvir J. Provocative tests with psychostimulant drugs in schizophrenia. Psychopharmacology (Berl) 1987;91:415–33. doi: 10.1007/BF00216006. [DOI] [PubMed] [Google Scholar]
  145. Lindstrom LH, Gefvert O, Hagberg G, Lundberg T. Increased dopamine synthesis rate in medial prefrontal cortex and striatum in schizophrenia indicated by L-(β- 11 C) DOPA and PET. Biol Psychiatry. 1999;46:681–688. doi: 10.1016/s0006-3223(99)00109-2. [DOI] [PubMed] [Google Scholar]
  146. Liu F, Guo X, Wu R, et al. Minocycline supplementation for treatment of negative symptoms in early-phase schizophrenia: a double blind, randomized, controlled trial. Schizophr Res. 2014;153:169–76. doi: 10.1016/j.schres.2014.01.011. [DOI] [PubMed] [Google Scholar]
  147. Lodge DJ, Grace AA. The hippocampus modulates dopamine neuron responsivity by regulating the intensity of phasic neuron activation. Neuropsychopharmacology. 2006;31:1356–1361. doi: 10.1038/sj.npp.1300963. [DOI] [PubMed] [Google Scholar]
  148. Lutkenhoff ES, van Erp TG, Thomas MA, et al. Proton MRS in twin pairs discordant for schizophrenia. Mol Psychiatry. 2010;15:308–318. doi: 10.1038/mp.2008.87. [DOI] [PubMed] [Google Scholar]
  149. Mackay aV, Iversen LL, Rossor M, et al. Increased brain dopamine and dopamine receptors in schizophrenia. Arch Gen Psychiatry. 1982;39:991–7. doi: 10.1001/archpsyc.1982.04290090001001. [DOI] [PubMed] [Google Scholar]
  150. Majo VJ, Prabhakaran J, Mann JJ, Kumar JSD. PET and SPECT tracers for glutamate receptors. Drug Discov Today. 2013;18:173–84. doi: 10.1016/j.drudis.2012.10.004. [DOI] [PubMed] [Google Scholar]
  151. Malaspina D, Harkavy-Friedman J, Corcoran C, et al. Resting neural activity distinguishes subgroups of schizophrenia patients. Biol Psychiatry. 2004;56:931–7. doi: 10.1016/j.biopsych.2004.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Malmberg a, Jackson DM, Eriksson a, Mohell N. Unique binding characteristics of antipsychotic agents interacting with human dopamine D2A, D2B, and D3 receptors. Mol Pharmacol. 1993;43:749–54. [PubMed] [Google Scholar]
  153. Manzoni O, Prezeau L, Desagher S, et al. Sodium nitroprusside blocks NMDA receptors via formation of ferrocyanide ions. Neuroreport. 1992;3 doi: 10.1097/00001756-199201000-00020. [DOI] [PubMed] [Google Scholar]
  154. Marsman A, van den Heuvel MP, Klomp DWJ, et al. Glutamate in schizophrenia: a focused review and meta-analysis of 1H-MRS studies. Schizophr Bull. 2013;39:120–9. doi: 10.1093/schbul/sbr069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Mason G, Petersen K, de Graaf RA, et al. Measurements of the anaplerotic rate in the human cerebral cortex using 13C magnetic resonance spectroscopy and [1-13 C] and [2-13C] glucose. J Neurochem. 2007;100:73–86. doi: 10.1111/j.1471-4159.2006.04200.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Matthysse S. Antipsychotic drug actions: a clue to the neuropathology of schizophrenia? Fed Proc. 1973;32:200–205. [PubMed] [Google Scholar]
  157. McCullumsmith R, Hammond J, Funk A, Meador-Woodruff JH. Recent advances in targeting the ionotropic glutamate receptors in treating schizophrenia. Curr Pharm Biotechnol. 2012;13:1535–1542. doi: 10.2174/138920112800784899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. McGinnity CJ, Hammers A, Riaño Barros Da, et al. Initial evaluation of 18F-GE-179, a putative PET Tracer for activated N-methyl D-aspartate receptors. J Nucl Med. 2014;55:423–30. doi: 10.2967/jnumed.113.130641. [DOI] [PubMed] [Google Scholar]
  159. Mcgowan S, Lawrence AD, Sales T. Presynaptic Dopaminergic Dysfunction in Schizophrenia. Arch Gen Psychiatry. 2004;61:134–142. doi: 10.1001/archpsyc.61.2.134. [DOI] [PubMed] [Google Scholar]
  160. McGuire P, Howes OD, Stone J, Fusar-Poli P. Functional neuroimaging in schizophrenia: diagnosis and drug discovery. Trends Pharmacol Sci. 2008;29:91–8. doi: 10.1016/j.tips.2007.11.005. [DOI] [PubMed] [Google Scholar]
  161. Meyer-Lindenberg A, Miletich RS, Kohn PD, et al. Reduced prefrontal activity predicts exaggerated striatal dopaminergic function in schizophrenia. Nat Neurosci. 2002;5:267–71. doi: 10.1038/nn804. [DOI] [PubMed] [Google Scholar]
  162. Miller DW, Abercrombie ED. Effects of MK-801 on spontaneous and amphetamine-stimulated dopamine release in striatum measured with in vivo microdialysis in awake rats. Brain Res Bull. 1996;40:57–62. doi: 10.1016/0361-9230(95)02144-2. [DOI] [PubMed] [Google Scholar]
  163. Miyake N, Skinbjerg M, Easwaramoorthy B, et al. Imaging changes in glutamate transmission in vivo with the metabotropic glutamate receptor 5 tracer [11C] ABP688 and N-acetylcysteine challenge. Biol Psychiatry. 2011;69:822–4. doi: 10.1016/j.biopsych.2010.12.023. [DOI] [PubMed] [Google Scholar]
  164. Miyaoka T, Yasukawa R. Minocycline as Adjunctive Therapy for Schizophrenia: An open-label study. Clin. 2008 doi: 10.1097/wnf.0b013e3181593d45. [DOI] [PubMed] [Google Scholar]
  165. Mizrahi R, Addington J, Rusjan PM, et al. Increased stress-induced dopamine release in psychosis. Biol Psychiatry. 2012;71:561–7. doi: 10.1016/j.biopsych.2011.10.009. [DOI] [PubMed] [Google Scholar]
  166. Mizrahi R, Kenk M, Suridjan I, et al. Stress-Induced Dopamine Response in Subjects at Clinical High Risk for Schizophrenia with and without Concurrent Cannabis Use. Neuropsychopharmacology. 2013;39:1479–1489. doi: 10.1038/npp.2013.347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Moghaddam B, Adams B, Verma A, Daly D. Activation of glutamatergic neurotransmission by ketamine: a novel step in the pathway from NMDA receptor blockade to dopaminergic and cognitive disruptions associated with the prefrontal cortex. J Neurosci. 1997;17:2921–2927. doi: 10.1523/JNEUROSCI.17-08-02921.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Monte AS, de Souza GC, McIntyre RS, et al. Prevention and reversal of ketamine-induced schizophrenia related behavior by minocycline in mice: Possible involvement of antioxidant and nitrergic pathways. J Psychopharmacol. 2013;27:1032–43. doi: 10.1177/0269881113503506. [DOI] [PubMed] [Google Scholar]
  169. Morgan CJ, Muetzelfeldt L, Curran HV. Ketamine use cognition and psychological wellbeing: a comparison of frequent, infrequent and ex-users with polydrug and non-using controls. Addiction. 2009;104:77–87. doi: 10.1111/j.1360-0443.2008.02394.x. [DOI] [PubMed] [Google Scholar]
  170. Morgan CJA, Curran HV. Acute and chronic effects of ketamine upon human memory: a review. Psychopharmacology (Berl) 2006;188:408–424. doi: 10.1007/s00213-006-0572-3. [DOI] [PubMed] [Google Scholar]
  171. Mortimer A, Singh P, Shepherd C, Puthiryackal J. Clozapine for Treatment-Resistant Schizophrenia: National Institute of Clinical Excellence (NICE) Guidance in the Real World. Clin Schizophr Relat Psychoses. 2010;4:49–55. doi: 10.3371/CSRP.4.1.4. [DOI] [PubMed] [Google Scholar]
  172. Murphy BP, Chung Y-C, Park T-W, McGorry PD. Pharmacological treatment of primary negative symptoms in schizophrenia: a systematic review. Schizophr Res. 2006;88:5–25. doi: 10.1016/j.schres.2006.07.002. [DOI] [PubMed] [Google Scholar]
  173. Murray M, Weinberger R, Herman M. Distribution of Putative D4 Dopamine Receptors Striatum from Patients with Schizophrenia in Postmortem. 1995;75 doi: 10.1523/JNEUROSCI.15-03-02186.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Narendran R, Frankle WG, Mason NS, et al. Positron emission tomography imaging of amphetamine-induced dopamine release in the human cortex: a comparative evaluation of the high affinity dopamine D2/3 radiotracers [11C]FLB 457 and [11C]fallypride. Synapse. 2009;63:447–61. doi: 10.1002/syn.20628. [DOI] [PubMed] [Google Scholar]
  175. Narendran R, Jedema HP, Lopresti BJ, et al. Imaging dopamine transmission in the frontal cortex: a simultaneous microdialysis and [11C]FLB 457 PET study. Mol Psychiatry. 2014;19:302–10. doi: 10.1038/mp.2013.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Natsubori T, Inoue H, Abe O, et al. Reduced Frontal Glutamate + Glutamine and N-Acetylaspartate Levels in Patients With Chronic Schizophrenia but not in Those at Clinical High Risk for Psychosis or With First-Episode Schizophrenia. Schizophr Bull. 2013 doi: 10.1093/schbul/sbt124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Nielsen MØ, Rostrup E, Wulff S, et al. Alterations of the brain reward system in antipsychotic naïve schizophrenia patients. Biol Psychiatry. 2012;71:898–905. doi: 10.1016/j.biopsych.2012.02.007. [DOI] [PubMed] [Google Scholar]
  178. Nordström AL, Farde L, Halldin C. Time course of D2-dopamine receptor occupancy examined by PET after single oral doses of haloperidol. Psychopharmacology (Berl) 1992;106:433–438. doi: 10.1007/BF02244811. [DOI] [PubMed] [Google Scholar]
  179. Nordstrom AL, Farde L, Wiesel FA, et al. Central D2-dopamine receptor occupancy in relation to antipsychotic drug effects: A double-blind PET study of schizophrenic patients. Biol Psychiatry. 1993;33:227–235. doi: 10.1016/0006-3223(93)90288-o. [DOI] [PubMed] [Google Scholar]
  180. Nozaki S, Kato M, Takano H, et al. Regional dopamine synthesis in patients with schizophrenia using L-[beta-11C]DOPA PET. Schizophr Res. 2009;108:78–84. doi: 10.1016/j.schres.2008.11.006. [DOI] [PubMed] [Google Scholar]
  181. Ohrmann P, Siegmund A, Suslow T, et al. Evidence for glutamatergic neuronal dysfunction in the prefrontal cortex in chronic but not in first-episode patients with schizophrenia: A proton magnetic resonance spectroscopy study. Schizophr Res. 2005;73:153–157. doi: 10.1016/j.schres.2004.08.021. [DOI] [PubMed] [Google Scholar]
  182. Okubo Y, Suhara T, Suzuki K, Kobayashi K. Decreased prefrontal dopamine D1 receptors in schizophrenia revealed by PET. Nature. 1997;385:634–636. doi: 10.1038/385634a0. [DOI] [PubMed] [Google Scholar]
  183. Olbrich HM, Valerius G, Rüsch N, et al. Frontolimbic glutamate alterations in first episode schizophrenia: evidence from a magnetic resonance spectroscopy study. World J Biol Psychiatry. 2008;9:59–63. doi: 10.1080/15622970701227811. [DOI] [PubMed] [Google Scholar]
  184. Olney JW, Farber NB. Glutamate receptor dysfunction and schizophrenia. Arch Gen Psychiatry. 1995;52:998–1007. doi: 10.1001/archpsyc.1995.03950240016004. [DOI] [PubMed] [Google Scholar]
  185. Ongur D, Prescot AP, Jensen JE, et al. Creatine abnormalities in schizophrenia and bipolar disorder. Psychiatry Res - Neuroimaging. 2009;172:44–48. doi: 10.1016/j.pscychresns.2008.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Owen F, Crow T, Poulter M, Cross A. Increased dopamine-receptor sensitivity in schizophrenia. Lancet. 1978;312:29–32. doi: 10.1016/s0140-6736(78)91740-3. [DOI] [PubMed] [Google Scholar]
  187. Patil ST, Zhang L, Martenyi F, et al. Activation of mGlu2/3 receptors as a new approach to treat schizophrenia: a randomized Phase 2 clinical trial. Nat Med. 2007;13:1102–1107. doi: 10.1038/nm1632. [DOI] [PubMed] [Google Scholar]
  188. Pearce R, Seeman P. Dopamine uptake sites and dopamine receptors in Parkinson’s disease and schizophrenia. Eur Neurol. 1990;30:9–14. doi: 10.1159/000117168. [DOI] [PubMed] [Google Scholar]
  189. Perreault ML, Hasbi A, Alijaniaram M, et al. The dopamine D1-D2 receptor heteromer localizes in dynorphin/enkephalin neurons: increased high affinity state following amphetamine and in schizophrenia. J Biol Chem. 2010;285:36625–34. doi: 10.1074/jbc.M110.159954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Perry TL. Normal cerebrospinal fluid and brain glutamate levels in schizophrenia do not support the hypothesis of glutamatergic neuronal dysfunction. Neurosci Lett. 1982;28:81–85. doi: 10.1016/0304-3940(82)90212-9. [DOI] [PubMed] [Google Scholar]
  191. Pilowsky LS, Bressan RA, Stone JM, et al. First in vivo evidence of an NMDA receptor deficit in medication-free schizophrenic patients. Mol Psychiatry. 2006;11:118–119. doi: 10.1038/sj.mp.4001751. [DOI] [PubMed] [Google Scholar]
  192. Pilowsky LS, Costa DC, Ell PJ, et al. Antipsychotic medication, D2 dopamine receptor blockade and clinical response: a 123I IBZM SPET (single photon emission tomography) study. Psychol Med. 1993;23:791–797. doi: 10.1017/s0033291700025575. [DOI] [PubMed] [Google Scholar]
  193. Pitcher G, Kalia L, Ng D, Goodfellow N. Schizophrenia susceptibility pathway neuregulin 1–ErbB4 suppresses Src upregulation of NMDA receptors. Nat Med. 2011;17:470–478. doi: 10.1038/nm.2315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Poels EMP, Kegeles LS, Kantrowitz JT, et al. Glutamatergic abnormalities in schizophrenia: A review of proton MRS findings. Schizophr Res. 2014;152:325–332. doi: 10.1016/j.schres.2013.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Pogarell O, Koch W, Karch S, et al. Dopaminergic neurotransmission in patients with schizophrenia in relation to positive and negative symptoms. Pharmacopsychiatry. 2012;45(Suppl 1):S36–41. doi: 10.1055/s-0032-1306313. [DOI] [PubMed] [Google Scholar]
  196. Reith J, Benkelfat C, Sherwin a, et al. Elevated dopa decarboxylase activity in living brain of patients with psychosis. Proc Natl Acad Sci U S A. 1994;91:11651–4. doi: 10.1073/pnas.91.24.11651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Remington G, Kapur ÞS, Foussias G, et al. Tetrabenazine Augmentation in Treatment-Resistant Schizophrenia. J Clin Pharmacol. 2012;32:95–99. doi: 10.1097/JCP.0b013e31823f913e. [DOI] [PubMed] [Google Scholar]
  198. Reynolds GP, Mason SL. Are striatal dopamine D4 receptors increased in schizophrenia? J Neurochem. 1994;63:1576–7. doi: 10.1046/j.1471-4159.1994.63041576.x. [DOI] [PubMed] [Google Scholar]
  199. Ripke S, Neale BM, Corvin A, et al. Biological insights from 108 schizophrenia-associated genetic loci. Nature. 2014 doi: 10.1038/nature13595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Roiser JP, Howes OD, Chaddock Ca, et al. Neural and behavioral correlates of aberrant salience in individuals at risk for psychosis. Schizophr Bull. 2013;39:1328–36. doi: 10.1093/schbul/sbs147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Roiser JP, Stephan KE, den Ouden HEM, et al. Do patients with schizophrenia exhibit aberrant salience? Psychol Med. 2009;39:199–209. doi: 10.1017/S0033291708003863. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Rosenheck R, Leslie D, Keefe R, et al. Barriers to employment for people with schizophrenia. Am J Psychiatry. 2006;163:411–7. doi: 10.1176/appi.ajp.163.3.411. [DOI] [PubMed] [Google Scholar]
  203. Rotaru DC, Lewis DA, Gonzalez-Burgos G. The role of glutamatergic inputs onto parvalbumin-positive interneurons: relevance for schizophrenia. Rev Neurosci. 2012;23:97–109. doi: 10.1515/revneuro-2011-0059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Rothman DL, Behar KL, Hyder F, Shulman RG. In vivo NMR studies of the glutamate neurotransmitter flux and neuroenergetics: implications for brain function. Annu Rev Physiol. 2003;65:401–427. doi: 10.1146/annurev.physiol.65.092101.142131. [DOI] [PubMed] [Google Scholar]
  205. Rowland LM, Spieker EA, Francis A, et al. White matter alterations in deficit schizophrenia. Neuropsychopharmacology. 2009;34:1514–1522. doi: 10.1038/npp.2008.207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Saha S, Chant D, Welham J, McGrath J. A systematic review of the prevalence of schizophrenia. PLoS Med. 2005;2:e141. doi: 10.1371/journal.pmed.0020141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Schlagenhauf F, Sterzer P, Schmack K, et al. Reward feedback alterations in unmedicated schizophrenia patients: relevance for delusions. Biol Psychiatry. 2009;65:1032–9. doi: 10.1016/j.biopsych.2008.12.016. [DOI] [PubMed] [Google Scholar]
  208. Schobel Sa, Chaudhury NH, Khan Ua, et al. Imaging patients with psychosis and a mouse model establishes a spreading pattern of hippocampal dysfunction and implicates glutamate as a driver. Neuron. 2013;78:81–93. doi: 10.1016/j.neuron.2013.02.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Seeman P. Dopamine agonist radioligand binds to both D2High and D2Low receptors, explaining why alterations in D2High are not detected in human brain scans. Synapse. 2012;66:88–93. doi: 10.1002/syn.20987. [DOI] [PubMed] [Google Scholar]
  210. Seeman P, Guan H, Van Tol H. dopamine d4 receptors elevated in schizophrenia. 1993 doi: 10.1038/365441a0. [DOI] [PubMed] [Google Scholar]
  211. Seeman P, Lee T. Antipsychotic drugs: direct correlation between clinical potency and presynaptic action on dopamine neurons. Science. 1975;188:1217–1219. doi: 10.1126/science.1145194. [DOI] [PubMed] [Google Scholar]
  212. Seeman P, Lee T, Chau-Wong M, Wong K. Antipsychotic drug doses and neuroleptic/dopamine receptors. Nature. 1976;261:717–719. doi: 10.1038/261717a0. [DOI] [PubMed] [Google Scholar]
  213. Seeman P, Niznik HB, Guan HC, et al. Link between D1 and D2 dopamine receptors is reduced in schizophrenia and Huntington diseased brain. Proc Natl Acad Sci U S A. 1989;86:10156–60. doi: 10.1073/pnas.86.24.10156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Seeman P, Schwarz J, Chen J-F et al. Psychosis pathways converge via D2high dopamine receptors. Synapse. 2006;60:319–346. doi: 10.1002/syn.20303. [DOI] [PubMed] [Google Scholar]
  215. Selvaraj S, Arnone D, Cappai A, Howes O. Alterations in the serotonin system in schizophrenia: A systematic review and meta-analysis of postmortem and molecular imaging studies. Neurosci Biobehav Rev. 2014;45C:233–245. doi: 10.1016/j.neubiorev.2014.06.005. [DOI] [PubMed] [Google Scholar]
  216. Sesack SR, Pickel VM. Prefrontal cortical efferents in the rat synapse on unlabeled neuronal targets of catecholamine terminals in the nucleus accumbens septi and on dopamine neurons in the ventral tegmental area. J Comp Neurol. 1992;320:145–60. doi: 10.1002/cne.903200202. [DOI] [PubMed] [Google Scholar]
  217. Sharp FR, Tomitaka M, Bernaudin M, Tomitaka S. Psychosis: Pathological activation of limbic thalamocortical circuits by psychomimetics and schizophrenia? Trends Neurosci. 2001;24:330–334. doi: 10.1016/s0166-2236(00)01817-8. [DOI] [PubMed] [Google Scholar]
  218. Shotbolt P, Stokes PR, Owens SF, et al. Striatal dopamine synthesis capacity in twins discordant for schizophrenia. Psychol Med. 2011;41:2331–8. doi: 10.1017/S0033291711000341. [DOI] [PubMed] [Google Scholar]
  219. Silvestri S, Seeman MV, Negrete J-C, et al. Increased dopamine D 2 receptor binding after long-term treatment with antipsychotics in humans: a clinical PET study. Psychopharmacology (Berl) 2000;152:174–180. doi: 10.1007/s002130000532. [DOI] [PubMed] [Google Scholar]
  220. Simpson EH, Kellendonk C, Kandel E. A possible role for the striatum in the pathogenesis of the cognitive symptoms of schizophrenia. Neuron. 2010;65:585–96. doi: 10.1016/j.neuron.2010.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Smith GS, Schloesser R, Brodie JD, et al. Glutamate modulation of dopamine measured in vivo with positron emission tomography (PET) and 11C-raclopride in normal human subjects. Neuropsychopharmacology. 1998;18:18–25. doi: 10.1016/S0893-133X(97)00092-4. [DOI] [PubMed] [Google Scholar]
  222. Snyder SH. The dopamine hypothesis of schizophrenia: focus on the dopamine receptor. Am J Psychiatry. 1976;133:197–202. doi: 10.1176/ajp.133.2.197. [DOI] [PubMed] [Google Scholar]
  223. Sokolov BP. Expression of NMDAR1, GluR1, GluR7, and KA1 glutamate receptor mRNAs is decreased in frontal cortex of “neuroleptic-free” schizophrenics: evidence on reversible up-regulation by typical neuroleptics. J Neurochem>. 1998;71:2454–64. doi: 10.1046/j.1471-4159.1998.71062454.x. [DOI] [PubMed] [Google Scholar]
  224. Soliman A, O’Driscoll Ga, Pruessner J, et al. Stress-induced dopamine release in humans at risk of psychosis: a [11C]raclopride PET study. Neuropsychopharmacology. 2008;33:2033–41. doi: 10.1038/sj.npp.1301597. [DOI] [PubMed] [Google Scholar]
  225. Sorce S, Schiavone S, Tucci P, et al. The NADPH oxidase NOX2 controls glutamate release: a novel mechanism involved in psychosis-like ketamine responses. J Neurosci. 2010;30:11317–11325. doi: 10.1523/JNEUROSCI.1491-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Stahl S. Stahl’s essential psychopharmacology: neuroscientific basis and practical applications. 2013:608. [Google Scholar]
  227. Stokes PRa, Shotbolt P, Mehta Ma, et al. Nature or nurture? Determining the heritability of human striatal dopamine function: an [18F]-DOPA PET study. Neuropsychopharmacology. 2013;38:485–91. doi: 10.1038/npp.2012.207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Stone JM. Glutamatergic antipsychotic drugs: a new dawn in the treatment of schizophrenia? Ther Adv Psychopharmacol. 2011;1:5–18. doi: 10.1177/2045125311400779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. Stone JM, Day F, Tsagaraki H. Glutamate Dysfunction in People with Prodromal Symptoms of Psychosis: Relationship to Gray Matter Volume. Biol Psychiatry. 2009;66:533–539. doi: 10.1016/j.biopsych.2009.05.006. [DOI] [PubMed] [Google Scholar]
  230. Stone JM, Howes OD, Egerton A, et al. Altered relationship between hippocampal glutamate levels and striatal dopamine function in subjects at ultra high risk of psychosis. Biol Psychiatry. 2010;68:599–602. doi: 10.1016/j.biopsych.2010.05.034. [DOI] [PubMed] [Google Scholar]
  231. Stone JM, Morrison PD, Pilowsky LS. Glutamate and dopamine dysregulation in schizophrenia--a synthesis and selective review. J Psychopharmacol. 2007;21:440–52. doi: 10.1177/0269881106073126. [DOI] [PubMed] [Google Scholar]
  232. Stone JM, Pepper F, Fam J, et al. Glutamate, N-acetyl aspartate and psychotic symptoms in chronic ketamine users. Psychopharmacology (Berl) 2014;231:2107–16. doi: 10.1007/s00213-013-3354-8. [DOI] [PubMed] [Google Scholar]
  233. Suridjan I, Rusjan P, Addington J, et al. Dopamine D2 and D3 binding in people at clinical high risk for schizophrenia, antipsychotic-naive patients and healthy controls while performing a cognitive task. J Psychiatry Neurosci. 2013;38:98–106. doi: 10.1503/jpn.110181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Szulc A, Konarzewska B, Galinska-Skok B, et al. Proton magnetic resonance spectroscopy measures related to short-term symptomatic outcome in chronic schizophrenia. Neurosci Lett. 2013;547:37–41. doi: 10.1016/j.neulet.2013.04.051. [DOI] [PubMed] [Google Scholar]
  235. Tamminga CA, Holcomb HH, Gao XM, Lahti AC. Glutamate pharmacology and the treatment of schizophrenia: current status and future directions. Int Clin Psychopharmacol. 1995;3(10 Suppl):29–37. [PubMed] [Google Scholar]
  236. Tandon N, Bolo NR, Sanghavi K, et al. Brain metabolite alterations in young adults at familial high risk for schizophrenia using proton magnetic resonance spectroscopy. Schizophr Res. 2013;148:59–66. doi: 10.1016/j.schres.2013.05.024. [DOI] [PubMed] [Google Scholar]
  237. Tayoshi S, Sumitani S, Taniguchi K, et al. Metabolite changes and gender differences in schizophrenia using 3-Tesla proton magnetic resonance spectroscopy (1H-MRS) Schizophr Res. 2009;108:69–77. doi: 10.1016/j.schres.2008.11.014. [DOI] [PubMed] [Google Scholar]
  238. Théberge J, Al-Semaan Y, Williamson PC, et al. Glutamate and glutamine in the anterior cingulate and thalamus of medicated patients with chronic schizophrenia and healthy comparison subjects measured with 4.0-T proton MRS. Am J Psychiatry. 2003;160:2231–2233. doi: 10.1176/appi.ajp.160.12.2231. [DOI] [PubMed] [Google Scholar]
  239. Théberge J, Bartha R, Drost DJ, et al. Glutamate and glutamine measured with 4.0 T proton MRS in never-treated patients with schizophrenia and healthy volunteers. Am J Psychiatry. 2002;159:1944–1946. doi: 10.1176/appi.ajp.159.11.1944. [DOI] [PubMed] [Google Scholar]
  240. Thompson JL, Urban N, Slifstein M, et al. Striatal dopamine release in schizophrenia comorbid with substance dependence. Mol Psychiatry. 2013;18:909–15. doi: 10.1038/mp.2012.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Tibbo P, Hanstock C, Valiakalayil A, Allen P. 3-T proton MRS investigation of glutamate and glutamine in adolescents at high genetic risk for schizophrenia. Am J Psychiatry. 2004;161:1116–1118. doi: 10.1176/appi.ajp.161.6.1116. [DOI] [PubMed] [Google Scholar]
  242. Tiihonen J, Halonen P, Wahlbeck K, et al. Topiramate add-on in treatment-resistant schizophrenia: a randomized, double-blind, placebo-controlled, crossover trial. J Clin Psychiatry. 2005;66:1012–1015. doi: 10.4088/jcp.v66n0808. [DOI] [PubMed] [Google Scholar]
  243. Tiihonen J, Wahlbeck K, Kiviniemi V. The efficacy of lamotrigine in clozapine-resistant schizophrenia: A systematic review and meta-analysis. Schizophr Res. 2009;109:10–14. doi: 10.1016/j.schres.2009.01.002. [DOI] [PubMed] [Google Scholar]
  244. Tsai GE, Lin P-Y. Strategies to enhance N-methyl-D-aspartate receptor-mediated neurotransmission in schizophrenia, a critical review and meta-analysis. Curr Pharm Des. 2010;16:522–537. doi: 10.2174/138161210790361452. [DOI] [PubMed] [Google Scholar]
  245. Tsukada H. [C] Raclopride Binding With No Alterations in Static Dopamine Concentrations in the Striatal Extracellular Fluid in the Monkey Brain : Multiparametric PET Studies Combined. 2000;103:95–103. doi: 10.1002/1098-2396(200008)37:2<95::AID-SYN3>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  246. Umbricht D, Alberati D, Martin-Facklam M, et al. Effect of Bitopertin, a Glycine Reuptake Inhibitor, on Negative Symptoms of Schizophrenia: A Randomized, Double-Blind, Proof-of-Concept Study. JAMA psychiatry. 2014 doi: 10.1001/jamapsychiatry.2014.163. [DOI] [PubMed] [Google Scholar]
  247. Vernaleken I, Klomp M, Moeller O, et al. Vulnerability to psychotogenic effects of ketamine is associated with elevated D2/3-receptor availability. Int J Neuropsychopharmacol. 2013;16:745–54. doi: 10.1017/S1461145712000764. [DOI] [PubMed] [Google Scholar]
  248. Videbaek C, Toska K, Scheideler Ma, et al. SPECT tracer [(123)I]IBZM has similar affinity to dopamine D2 and D3 receptors. Synapse. 2000;38:338–42. doi: 10.1002/1098-2396(20001201)38:3<338::AID-SYN13>3.0.CO;2-N. [DOI] [PubMed] [Google Scholar]
  249. Volk S, Maul FD, Hör G, et al. Dopamine D2 receptor occupancy measured by single photon emission computed tomography with 123I-Iodobenzamide in chronic schizophrenia. Psychiatry Res. 1994;55:111–118. doi: 10.1016/0925-4927(94)90005-1. [DOI] [PubMed] [Google Scholar]
  250. Vollenweider FX, Vontobel P, Oye I, et al. Effects of (S)-ketamine on striatal dopamine: A [11C]raclopride PET study of a model psychosis in humans. J Psychiatr Res. 2000;34:35–43. doi: 10.1016/s0022-3956(99)00031-x. [DOI] [PubMed] [Google Scholar]
  251. Walinder J, Skott A. Potentiation by Metyrosine of Thioridazine Effects in Chronic Schizophrenics a Long-Term Trial Using Double-Blind Crossover Technique. Arch Gen Psychiatry. 1976;33:501–505. doi: 10.1001/archpsyc.1976.01770040061011. [DOI] [PubMed] [Google Scholar]
  252. Wang M, Pei L, Fletcher PJ, et al. Schizophrenia, amphetamine-induced sensitized state and acute amphetamine exposure all show a common alteration: increased dopamine D2 receptor dimerization. Mol Brain. 2010;3:25. doi: 10.1186/1756-6606-3-25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  253. Ward RD, Simpson EH, Richards VL, et al. Dissociation of hedonic reaction to reward and incentive motivation in an animal model of the negative symptoms of schizophrenia. Neuropsychopharmacology. 2012;37:1699–707. doi: 10.1038/npp.2012.15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  254. Weinberger D. Implications of normal brain development for the pathogenesis of schizophrenia. Arch Gen Psychiatry. 1987;45:1055. doi: 10.1001/archpsyc.1987.01800190080012. [DOI] [PubMed] [Google Scholar]
  255. Weiser M, Heresco-Levy U, Davidson M, et al. A multicenter, add-on randomized controlled trial of low-dose d-serine for negative and cognitive symptoms of schizophrenia. J Clin Psychiatry. 2012;73:e728–34. doi: 10.4088/JCP.11m07031. [DOI] [PubMed] [Google Scholar]
  256. Whiteford HA, Degenhardt L, Rehm J, et al. Global burden of disease attributable to mental and substance use disorders: findings from the Global Burden of Disease Study 2010. Lancet. 2013;382:1575–86. doi: 10.1016/S0140-6736(13)61611-6. [DOI] [PubMed] [Google Scholar]
  257. Winton-Brown T, Fusar-Poli P, Ungless M, Howes O. Dopaminergic basis of salience dysregulation in psychosis. Trends Neurosci. 2014;37:85–94. doi: 10.1016/j.tins.2013.11.003. [DOI] [PubMed] [Google Scholar]
  258. Wolkin A. Dopamine Receptor Occupancy and Plasma Haloperidol Levels. Arch Gen Psychiatry. 1989;46:482. doi: 10.1001/archpsyc.1989.01810050096021. [DOI] [PubMed] [Google Scholar]
  259. Zhang L, Shirayama Y, Iyo M, Hashimoto K. Minocycline attenuates hyperlocomotion and prepulse inhibition deficits in mice after administration of the NMDA receptor antagonist dizocilpine. Neuropsychopharmacology. 2007;32:2004–2010. doi: 10.1038/sj.npp.1301313. [DOI] [PubMed] [Google Scholar]
  260. Zwanzger P, Zavorotnyy M, Gencheva E, et al. Acute shift in glutamate concentrations following experimentally induced panic with cholecystokinin tetrapeptide--a 3T-MRS study in healthy subjects. Neuropsychopharmacology. 2013;38:1648–54. doi: 10.1038/npp.2013.61. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES