Acessibilidade / Reportar erro

Translational neuropsychiatry of genetic and neurodevelopmental animal models of schizophrenia

Abstracts

Psychiatric symptoms are subjective by nature and tend to overlap between different disorders. The modelling of a neuropsychiatric disorder therefore faces challenges because of missing knowledge of the fundamental pathophysiology and a lack of accurate diagnostics. Animal models are used to test hypotheses of aetiology and to represent the human condition as close as possible to increase our understanding of the disease and to evaluate new targets for drug discovery. In this review, genetic and neurodevelopmental animal models of schizophrenia are discussed with respect to behavioural and neurophysiological findings and their association with the clinical condition. Only specific animal models of schizophrenia may ultimately lead to novel diagnostic approaches and drug discovery. We argue that molecular biomarkers are important to improve animal to human translation since behavioural readouts lack the necessary specificity and reliability to assess human psychiatric symptoms.

Translational research; neuropsychiatry; genetic; neurodevelopment; animal model; schizophrenia


Sintomas psiquiátricos são subjetivos por natureza e tendem a se sobrepor entre diferentes desordens. Sendo assim, a criação de modelos de uma desordem neuropsiquiátrica encontra desafios pela falta de conhecimento dos fundamentos da fisiopatologia e diagnósticos precisos. Modelos animais são usados para testar hipóteses de etiologia e para representar a condição humana tão próximo quanto possível para aumentar nosso entendimento da doença e avaliar novos alvos para a descoberta de drogas. Nesta revisão, modelos animais genéticos e de neurodesenvolvimento de esquizofrenia são discutidos com respeito a achados comportamentais e neurofisiológicos e sua associação com a condição clínica. Somente modelos animais específicos de esquizofrenia podem, em último caso, levar a novas abordagens diagnósticas e descoberta de drogas. Argumentamos que biomarcadores moleculares são importantes para aumentar a tradução de animais a humanos, já que faltam a especificidade e a fidelidade necessárias às leituras comportamentais para avaliar sintomas psiquiátricos humanos.

Pesquisa traducional; neuropsiquiatria; genética; neurodesenvolvimento; modelo animal; esquizofrenia


Translational neuropsychiatry of genetic and neurodevelopmental animal models of schizophrenia

Michael G. GottschalkI; Zóltan SarnyaiII,III; Paul C. GuestI; Laura W. HarrisI; Sabine BahnI,IV

IDepartment of Chemical Engineering and Biotechnology, University of Cambridge, Cambridge, United Kingdom

IIDepartment of Pharmacology, University of Cambridge, Cambridge, United Kingdom

IIIDiscipline of Physiology & Pharmacology, James Cook University, Townsville, Australia

IVDepartment of Neuroscience, Erasmus Medical Centre, Rotterdam, The Netherlands

Address correspondence to

ABSTRACT

Psychiatric symptoms are subjective by nature and tend to overlap between different disorders. The modelling of a neuropsychiatric disorder therefore faces challenges because of missing knowledge of the fundamental pathophysiology and a lack of accurate diagnostics. Animal models are used to test hypotheses of aetiology and to represent the human condition as close as possible to increase our understanding of the disease and to evaluate new targets for drug discovery. In this review, genetic and neurodevelopmental animal models of schizophrenia are discussed with respect to behavioural and neurophysiological findings and their association with the clinical condition. Only specific animal models of schizophrenia may ultimately lead to novel diagnostic approaches and drug discovery. We argue that molecular biomarkers are important to improve animal to human translation since behavioural readouts lack the necessary specificity and reliability to assess human psychiatric symptoms.

Keywords: Translational research, neuropsychiatry, genetic, neurodevelopment, animal model, schizophrenia.

Introduction

According to a prediction of the World Health Organization (WHO) neuropsychiatric disorders including schizophrenia, depression, bipolar disorder and dementia, will be the greatest factor contributing to disease-adjusted life years by the year 20301. Epidemiological studies have calculated a lifetime morbidity risk for schizophrenia of approximately 1%2. Fundamental research into the psychopathology of this disorder and drug development during recent years has yielded few answers and more questions. One of the major reasons for this lack of progress appears to result from the diversity of psychiatric disorders, which often overlap in subjectively evaluated symptoms.

Clinical research calls for animal models as a necessary first step towards a deeper understanding of psychopathology and to provide tools for successful drug discovery. However, current models are only able to mimic certain psychiatric symptoms and it is questionable if we will ever be able to mirror and evaluate a multifaceted disorder such as schizophrenia using animal models. To this end, three categories are now assessed for animal models: construct, face and predictive validity3.

Construct validity

The degree to which animal models embody and conform to etiological aspects of the disease is termed the construct validity. One difficulty with the study of schizophrenia is that we cannot simply knock out one risk gene in order to generate symptoms. This is because schizophrenia, like most neuropsychiatric disorders, it is a multi-locus and multi-factorial disorder. Also, the precipitating environmental factors are either unknown or cause a range of psychiatric syndromes. This lack of certainty impedes creation of animal models with features specific for one psychiatric disorder with high construct validity.

Face validity

Face validity is the extent to which animal models reflect the main features or symptoms associated with a psychiatric disorder. This has most commonly been assessed using behavioural testing. A major barrier to development of animal models with high face validity is that the primary symptoms of schizophrenia, such as hallucinations, delusions, speech abnormalities and inappropriate behaviour, are impossible to model in other species. Nonetheless, several behavioural paradigms have been developed that have some relevance to schizophrenia, including hyperlocomotion, social interaction deficits, effects on prepulse inhibition (PPI), and intra- and extra-dimensional (ID/ED) shifting. The controversial association of hyperlocomotion with schizophrenia positive symptoms is based on its post-pubertal onset and exacerbation with N-methyl-D-aspartate (NMDA) receptor antagonists, indicating subcortical dopamine excess4. Social interaction deficits are associated with the negative symptom of social withdrawal in patients which has been shown to have predictive power for the development of schizophrenia in children5-7.

The Measurement and Treatment Research to Improve Cognition in Schizophrenia (MATRICS) identified seven domains of cognitive dysfunctions in schizophrenia patients including specific tests for measurement of these: working memory, verbal memory, visual learning, attention/vigilance, abstract reasoning, social cognition and speed of processing8-10. Unlike the changing intensity of psychosis, cognitive symptoms tend to accumulate and treatment of these has been found to be a better predictor of therapeutic efficiency and patient outcome than treatment of positive or negative symptoms11. Learning and memory tasks are translatable to rodents and numerous behavioural paradigms are available. Sensorimotor gating deficits are considered endophenotypes of schizophrenia, belonging to the category of cognitive impairments. Such deficits are related to the inability of schizophrenia patients to ignore irrelevant stimuli, and are believed to mirror the positive symptoms causing learning deficits and cognitive fragmentation in schizophrenia12-14. Human gating defects include reduced P50 auditory evoked potential suppression and inhibition of the startle reflex (both measurements of PPI), and abnormal smooth eye pursuit, which may reflect fundamental causes for the clinical expression of schizophrenia15,16. An advantage of such tests is the homology between the human and the animal test conditions17. More recently, a rodent analogue of the Wisconsin card sorting test (attentional set shifting task) has been developed18. Performance in this task measures prefrontal function and is impaired in schizophrenia patients.

Although all of these behavioural hallmarks have translational features, they are not likely to be disease-specific19-21. Furthermore, all existing animal models represent only distinct features of the disease. Considering the complexity of the disorder, it remains questionable whether we will achieve creation of an ideal model. Therefore, there is an urgent need for identification of translatable molecular markers in order to increase confidence in the validity of available animal models.

Predictive validity

The utility of a model in the prediction of drug efficacy in the clinic is termed the predictive validity. Assessment of this aspect is compromised by our limited knowledge of drug effects. Since the first antipsychotic drugs were discovered serendipitously based on dopamine receptor blockade, all new drugs have been developed in models chosen for their predictive validity as dopamine antagonists. However, recent studies have demonstrated the influence of antipsychotics on multiple neurotransmitter systems and signalling cascades, and progress has been made in moving model development away from dopamine pathways.

Aim of this review

This review focuses on relevant behavioural and neurophysiological findings in currently used animal models of schizophrenia from the perspective of translational neuropsychiatry. Our goal is not to list all models but to focus on differential etiological theories of schizophrenia and on the models which target these pathways. Following the assumptions of the Research Domain Criteria (RDoC) project launched by the NIMH, clinical neuroscience research will yield biomarkers to extend the view on psychiatric symptoms and improve patient management22. We believe that the correlation of schizophrenic patients and psychiatric animal models will result in a greater insight in psychopathology and provide the research tools essential for novel drug discovery strategies.

Reciprocity of research: finding the "right" aetiology

Many different views on the aetiology of schizophrenia have emerged and it is now believed that the pathogenesis results from an interaction of genetic and environmental risk factors3. The neuropathological features of the disorder include dysregulation of neurotransmitter systems (primarily dopamine and glutamate)23,24, hypofrontality25 and hippocampal synaptic deficits26. All of these have been followed by construction of relevant animal models resulting in a range of findings. The aetiological validity of an animal model can only be considered as a measurement of the precision of our knowledge of human pathogenesis27,28. In spite of the great diversity of aetiological findings in schizophrenia research, some aspects have been more reproducible in epidemiological studies. This includes genetic heritability and vulnerability during neurodevelopment which form the main focus of this review (Figure 1)29-31.


Gift of our ancestors: genes

Genetically modified rodents have become important research models in medical sciences in recent years. Modern molecular methods allow for targeted gene approaches and can help in discovering the neurobiological effects of genes and their interplay on behaviour. Genome-Wide Association Studies (GWAS) have facilitated identification of risk factors including single nucleotide polymorphisms, copy number variations and structural or allelic variants, and these can be mirrored in animal models. The heritability of schizophrenia has been estimated to be around 73% to 90%, based on twin studies and currently around 30 risk alleles have been identified32-34. Among the polymorphism- and chromosomal-based schizophrenia mouse models are the phenotypically-well characterized disrupted in schizophrenia 1 (DISC1), neuregulin 1 (NRG1), proline dehydrogenase (PRODH) and 22q11 mutants35. Further models are represented by dysbindin (DTNBP1) and reelin (RELN) knock-outs36,37. As with all genetic models, the resulting experimental findings should be considered with care since schizophrenia appears to be an oligogenetic disease with multiple small risk genes accumulating to a give a particular phenotype38-40.

DISC1

DISC1 was first discovered due to its co-segregation with psychiatric disorders in a Scottish family and it has been found to be an important vulnerability factor for schizophrenia and other mental disorders in diverse populations41,42. DISC1 has been shown to play an essential role in pre- and post-natal neurodevelopment, particularly in synaptogenesis, synaptic plasticity and neuronal migration43. The commonly used 129S6 Sv/Ev mouse strain has been shown to carry a 25bp deletion in exon 6 of this gene, which may result in a working memory deficit42. Disorganization and impaired plasticity of neurons may be the cause of the cognitive deficits found in DISC1 mice26. The N-ethyl-N-nitrosourea (ENU)-induced Disc1 missense mutation, L100P, produced schizophrenia-like behavioural impairments in PPI, latent inhibition and hyperlocomotion which could be blocked by administration of typical and atypical antipsychotics44. Working memory deficits were also reported for a mouse model of complete DISC1 deletion45. The Q31L missense mutation resulted in the generation of mice with deficits in working memory and diminished social interaction, and this could be blocked with clozapine44. Both mutant strains showed reduced volumes of different brain structures38,46,47. Consistent with these findings, DISC1 has been shown to be developmentally regulated47,48. At the molecular level, reduced DISC1 binding to the cAMP-specific 3',5'-cyclic phosphodiesterase 4B (PDE4B) has been demonstrated. PDE4B is a phosphodiesterase which has been implicated in animal models with learning and memory impairments44,49,50. Polymorphisms of PDE4B have been reported to be associated with schizophrenia in human GWAS51. Approaches with mutated human DISC1 in mice have resulted in reports of hyperlocomotion, diminished reward learning, impairments in social interaction and declarative memory and increased aggression52-54. Structural abnormalities in this model have been reported such as enlarged ventricles and decreased dendritic arborisation, with decreased expression of lissencephaly 1 (LIS1) and synaptosomal-associated protein 25 (SNAP-25)53. LIS1, DISC1 and nuclear distribution protein nudE-like 1 (NDEL1) have been shown to be part of a molecular interaction complex relevant in neurodevelopment55-58. Morphological changes in DISC1 mice have been observed including reduced prefrontal cortex volume and impaired growth and dendritic orientation in the dentate gyrus, correlating with findings in human schizophrenia studies59-61.

NRG1

NRG1 has been implicated by GWAS as a risk factor of schizophrenia62. The roles of NRG1 and its various isoforms include synaptogenesis, myelination, glia-neuron interactions, glial cell formation, NMDA receptor-regulated neuronal migration and glutamatergic signalling at a range of neurodevelopmental stages63-65. Behavioural abnormalities have been observed in mice lacking one of the NRG1 isoforms, including reduced fear conditioning and lower locomotor activity66,67. Some of these effects have been shown to be reversed by administration of atypical antipsychotics68. Behavioural analyses of NRG knockout mice have shown decreased sensorimotor gating and social interaction, with increased aggression69-72. Molecular analyses of the NMDA receptor in NRG1 knockout mice revealed an overall decrease and hypophosphorylation which could be reversed by clozapine treatment62,73.

22q11

In schizophrenia research, several studies have been performed on the chromosomal microdeletion at 22q11.2 (22q11DS), which results in velocardiofacial syndrome (VCFS), with palatal abnormalities, heart defects, microcephaly and deformation of thymus or parathyroid glands. Some of the cognitive abnormalities observed in patients with 22q11DS are thought to originate in the prefrontal cortex and hippocampus including deficits in attention, working memory, executive function and short-term verbal memory74-76. Approximately 25% of 22q11DS patients suffer from schizophrenia or schizoaffective disorder77-79. Also, 6% of early onset schizophrenia patients have the same microdeletion80,81. Several genes in this chromosomal region have been associated with schizophrenia including proline dehydrogenase (PRODH), catechol O-methyl transferase (COMT) and guanine nucleotide-binding protein subunit beta-like protein 1 (GNB1L)82-84. The COMT enzyme has been reported to modulate degradation of dopamine in the brain and therefore its absence could be responsible for working memory deficits75. Mutations of PRODH have been shown to produce neurotransmitter abnormalities in glutamatergic and dopaminergic pathways of genetically modified mice75. Selective deletion of PRODH in mice has been shown to lead to deficits in sensorimotor gating, with decreased glutamate, aspartate and GABA levels in the frontal cortex85. Proline has been implicated in modulation of neurotransmission especially in the glutamatergic system, which could lead to the psychotic symptoms observed in patients86-88. 22q11DS mouse models present with a loss of glutamatergic synapses, decreased spine density, decreased CA1 pyramidal neuron complexity and reduced neurogenesis in the subventricular zone. The same mice show behavioural abnormalities in PPI, hyperlocomotion, fear conditioning and working memory89-94. 22q11DS is the only confirmed recurrent structural mutation associated with schizophrenia identified to date95.

Mutations in neuronal signal transmission pathways

Neurotransmitter deficits have been modelled using genetic methods. Higher dopamine concentrations in the synaptic cleft can be simulated by knock-out of the dopamine transporter (DAT). Such mice have been classified as models of schizophrenia as they are hyperactive in novel environments, express excessive stereotypic behaviour, reduced reward learning, deficits in sensorimotor gating, working and declarative memory, and show circadian rhythm impairments96-99. The modified neuron-oscillatory properties caused by DAT knock-out have also been shown to influence the serotonergic system, possibly as compensation for the hyperdopaminergia100,101. The hypothalamic-pituitary-adrenal axis in these mice is dysregulated, resulting in anterior pituitary hypoplasia102, similar to effects seen in schizophrenic patients103. Also, approximately 50% lower concentrations of D1 and D2 postsynaptic receptor mRNA have been described in the striatum of these animals as well as downregulation of preproenkephalin A (PENK) and upregulation of dynorphin (PDYN)96,97. Following the glutamate hypothesis, a mutant mouse strain has been created with a 90% reduction in NMDA receptor 1 (NR1) expression, resulting in social and sexual behavioural deficits and similar symptoms to the DAT knock out mouse. These include impaired circadian rhythm, memory and gating deficits, diminished reward learning and increased locomotion104-107. These symptoms could be reduced with haloperidol and clozapine treatment104. Reduced NMDA-signalling induced by gene knockdown resulted in impaired habituation of the auditory startle response, similar to that induced by phencyclidine (PCP) administration, and mimicking the sensory overload believed to be experienced by many schizophrenic patients106,108. The observed loss of parvalbumin-positive neurons in NMDA deficient mice has also been linked with schizophrenia-like symptoms and has been reported for hippocampal interneurons21,109,110.

Vulnerability of our youth: neurodevelopment

The existence of an environmental component of schizophrenia was demonstrated by the finding that identical twins show a concordance of approximately 50%111. The neurodevelopmental character of schizophrenia was postulated based on epidemiological, human post-mortem tissue and imaging studies112. Neurodevelopmental animal models include maternal immune activation, prenatal malnutrition, the interruption of neurogenesis by methylazoxymethanol (MAM) and the neonatal ibotenic acid lesion of the ventral hippocampus113,114. Additional models are represented by the exposure of pregnant mice to unpredictable stress, maternal separation or birth complications115-117. All of these models have revealed behavioural deficits associated either with positive, cognitive or negative symptoms of schizophrenia, and alterations in the dopaminergic, glutamatergic, GABAergic and serotonergic systems118-124.

Maternal immune activation

Maternal immune activation is considered to be an etiological factor in schizophrenia based on epidemiological studies of influenza infection in mothers of schizophrenia patients, and an increased incidence of schizophrenia among winter births125-127. This has been modelled in rodents by infecting pregnant animals with viruses or by administration of the bacterial endotoxin lipopolysaccharide (LPS). Sensorimotor gating was reported to be impaired in male and female offspring of mothers treated with LPS between gestation day 15 and 19, but the effects in male mice were more severe128. A reduction of glycogen synthase kinase 3 β (GSK3B) mRNA in the prefrontal cortex was consistent with the findings of human post-mortem tissue analyses128-131. Male offspring of maternal immune activation showed an increase in tumour necrosis factor α (TNF) throughout their lifespan, as seen in schizophrenia128,132,133. Antipsychotics have been shown to reverse the effects of maternal immune activation on the behavioural and serum cytokine levels134,135. Maternal infection of rodents with influenza virus has led to deficits in social interaction and acoustic PPI in the offspring136. Prenatal influenza infection has been associated with impaired corticogenesis and cell layer disruptions in the hippocampus, together with altered expression of synaptic markers118,137. This is similar to the finding of hippocampal pyramidal cell disarray reported in human schizophrenia post-mortem studies138-140. Because the influenza virus or influenza virus antibodies have not been detected in the fetal brain, it has been hypothesized that the maternal immune response alone is sufficient to generate schizophrenia-like behavioural findings in the offspring136,141-145. At the neurotransmitter level, dysfunctions in the dopaminergic and glutamatergic system have been reported, especially in the medial prefrontal cortex and the nucleus accumbens in offspring from mothers subjected to immune challenge123.

Maternal protein malnutrition

Prenatal malnutrition has been shown to be a risk factor for schizophrenia146-148. Maternal malnutrition in rats was followed by attention and gating deficits on the behavioural level but only in female offspring149. This was associated with increased 3H-Haloperidol binding in the striatum, as found in studies of male rats following maternal malnutrition121,149, although there are conflicting reports of this150. Increased 3H-MK-801 binding in the striatum and hippocampus was described in the maternal malnutrition model and may reflect a NMDA receptor hypofunction and subsequent up-regulation as postulated by the glutamate hypothesis of schizophrenia121,151. Memory deficits in this model were described in terms of susceptibility to interference and extinction, further supporting theories of prefrontal and hippocampal psychopathology152. Offspring of malnourished mothers showed reduced social interaction and higher aggressive behaviour, similar to findings in malnourished children153,154. Increased concentrations of noradrenaline and dopamine were reported following malnutrition, resulting most likely from tyrosine hydroxylase hyperactivity and this correlated with reduced overall brain and hippocampus size155. Additionally GABAA receptors (GABAAR) have been reported to be decreased in the hippocampus of adult offspring subjected to maternal malnutrition119.

Interruption of neurogenesis

Administration of the mitotoxin methylazoxymethanol acetate (MAM) to rat dams on embryonic day 17 (E17) interferes with development of schizophrenia-relevant brain regions in the offspring114,120,156,157. Several mechanisms of action have been proposed to explain the actions of MAM-E17 treatment, including a decrease in the levels of reelin and the 67 kDa form of glutamic acid decarboxylase (GAD67), as reported in schizophrenic patients, or by methylation of genes critical for neuronal development and plasticity158-160. Neuroanatomical findings reported in parallel for the MAM-E17 model and schizophrenic patients include morphological changes in the prefrontal cortex, parahippocampal cortex and hippocampus, reduced size of the medial dorsal thalamus and increased neuron packaging in the frontal lobe114,161-165. A decrease in whole brain volume of MAM-E17 rats was mainly focussed on the prefrontal cortex, hippocampus and nucleus accumbens120. Cellular disarray in the CA1 region of hippocampus has also been described and is thought to be analogous to hippocampal disarray seen in schizophrenia patients120,166. Recently, we carried out a proteomic and metabonomic investigation which showed that MAM-E17 rats have primarily deficits in hippocampal glutamatergic neurotransmission, as seen in some schizophrenia patients167. Most importantly, these results were consistent with our finding of functional deficits in glutamatergic neurotransmission, as identified using electrophysiological recordings. At the behavioural level, dysfunctions in sensorimotor gating in MAM-E17 rats have been described in line with findings in schizophrenic patients, implicating frontal dysfunctions and dopaminergic input to the striatum114,168. Social interaction deficits have been described to occur before and after puberty in MAM-E17 rats, although hypermotor activity and deficits in spatial working memory and PPI occurred only after puberty169. Furthermore, the offspring of MAM-E17 rats displayed working memory deficits in object recognition, as do schizophrenic patients, implicating deficits in the prefrontal cortex120,170, 171. Additional evidence for prefrontal abnormalities has been reported by impairments in latent inhibition120,172-174. Cortical pyramidal neurons from MAM-E7 offspring failed to reduce their firing frequency as did control neurons if dopamine was artificially injected175. This finding has relevance to conditional associative learning which has been described as being impaired in schizophrenia patients176.

Neonatal hippocampal lesion

The neonatal hippocampal lesion rat model was developed after reports that this region affected mesolimbic dopamine transmission and cortical dopamine turnover in schizophrenia113. Ibotenic acid was administered on postnatal day 7 to damage the ventral hippocampus, which connects to the prefrontal cortex, analogous to the human anterior hippocampus. These animals showed hypermotor activity and hyper-responsiveness to stress emerging in adulthood (day 56). Also, the effect could be blocked by haloperidol treatment113. Conversely, sensorimotor gating deficits in this model have been shown to be treatable with the atypical antipsychotics clozapine, olanzapine and risperidone but not with the typical antipsychotic haloperidol177. This may be a useful model for schizophrenia due to the post-pubertal emergence of PPI deficits, as seen in the human disease, which are thought to result from damage to the prefrontal cortex17,178,179. Further behavioural tests have revealed deficits in spatial learning and memory, beginning in the juvenile period180. Reduced social interaction has been reported in this model prior to and after puberty, although this effect does not appear to respond to clozapine treatment181. This may be relevant in light of studies showing that prodromal individuals demonstrate difficulties in establishing friendships and emotional bonding5,182-184. Molecular studies have shown increased AMPA receptor (GRIA3) levels in the prefrontal cortex of ventral hippocampal lesion rats, which may be the cause of striatal dopaminergic hyperactivation185,186. Finally, studies on neuronal circuitry in the prefrontal cortex in this model have shown abnormal synaptic responses to excitatory input, which is of interest given the deficit in inhibitory interneurons found in the prefrontal cortex of schizophrenic post-mortem tissue124,187,188.

Translation of animals to humans and back again

Currently, animal behaviours which parallel specific characteristics of psychiatric disorders such as schizophrenia are used in drug development as early indicators of potential treatment efficacy. However, the use of behavioural readouts is limited in disease specificity and this has been regarded as one of the major drawbacks of these approaches189. In particular, the assessment of similarities of animal behaviour and patient characteristics can be subjective, leading to problems of bias and irreproducibility. Due to such difficulties, the disease relevance of many animal models for schizophrenia remains dubious.

One means of overcoming these problems is the demonstration that data accumulated from relevant animal model studies are re­presentative of human schizophrenia190. A simple approach would be to consider the molecular changes in blood serum or plasma which have been associated consistently with schizophrenia and attempt to correlate these findings with those in animal models of schizophrenia which target particular aspects of the disease. For example similar changes have been identified between specific animal models and human schizophrenia in the case of molecules such as adrenocorticotrophic hormone191, corticosteroid hormones192, brain-derived neurotrophic factor192, interleukin 2193, interleukin 6193, kynurenic acid194 and glutamate195. Such parallels in molecular changes may lead to the establishment of construct validity of a given animal model.

The efficient use of reverse translational research seems likely to guide the way for characterization of disease mechanisms and identifying novel drug targets. This is critical as the future clinical need in the field of neuropsychiatry calls for identification of class- or even drug-specific biomarkers matched with novel therapeutic approaches. This could be done by statistical analyses of the complex data generated in proteomic approaches in animals and humans alike. Once discovered and matched, new biomarkers could be used to evaluate the comparability of the animal model and the disease status in humans and treatment effects could be validated more efficiently than with behavioural tests. The goal of these molecular profiling approaches is the identification of cross species signatures of a given human disease. In this way, the best animal models for developing and testing novel therapeutic strategies can be identified.

Conclusions

This review summarized behavioural findings and key features on the morphological, cellular and molecular level in commonly used animal models of schizophrenia generated with genetic and neurodevelopmental approaches. It should be no surprise that none of the listed animal models has been established as the "best" for use in drug discovery, because none are able to mimic the human disease state in all measurable variables. Psychiatric patients present with multiple subtle symptoms, which cannot be measured reproducibly by existing techniques. Therefore, it is not likely that such symptoms can be recapitulated in animal models. In fact, several observations have been made which negatively impact on face validity such as the appearance of higher locomotor activity in some models which has no distinct counterpart in schizophrenic patients196. The need for this paradigm change is reflected in the ongoing debate about the introduction of the Diagnostic and Statistical Manual V (DSM-V). Current psychiatric practice includes the assessment of patient symptoms using clinically approved interviews which are not based on the underlying pathophysiology. It should be no surprise that clinical syndromes based on subjective symptoms are a source of misdiagnosis and therefore mistreatments.

To address the problems of combining the next generation of diagnostic rules with the findings of modern neuroscience, the National Institute of Mental Health has initiated the establishment of a classification guide of psychiatric patients for research proposes and for a better incorporation of modern scientific knowledge on the field of neuropsychiatry22. The RDoC provides a multidimensional view of "constructs" like behaviourally evaluated cognitive deficits and molecular neurobiological alterations. The major advantage of this approach is the freedom of diagnoses for certain syndromes, circumventing the problem of disease categories. RDoC represents a dimensional system ranging from normality to abnormality. With the exploitation of physiological measures and molecular biomarkers, this will enable a better integration of animal models in pre-clinical research and enhance their potential to generate value from bench to bedside. Recently new approaches have emerged to integrate new parameters into this process, such as molecular changes in easy accessible peripheral samples (for example saliva) and electrocardiographic measurements for early detection and diagnosis of stress-related psychopathogenesis in humans and for verification of animal models of psychiatric disorders197.

This review has focussed on the genetic and neurodevelopmental models of schizophrenia. Despite the fact that these were created to model schizophrenia, they represent great variability in their behavioural findings and it is difficult to evaluate them with respect to given symptoms. In the case of human psychiatric disorders, symptoms often overlap between different psychiatric disorders and psychiatrists are faced with a diversity of possible differential diagnoses (Figure 1). Dysfunctions of PPI, for example, have been reported in many neuropsychiatric and even in some non-neuropsychiatric disorders198.

Molecular findings have been used to show similarities between humans and rodents supporting the view that behaviour alterations are linked to molecular changes190,199. Given the reported pool of findings it seems unlikely that a difference in any single molecule would be enough to assign a given animal model to a distinct psychiatric disorder. This is in line with biomarker findings which have focussed on altered patterns rather than changes of single proteins200. For example it has been shown that a combination of 51 serum biomarkers can be used in humans to help with the diagnosis of schizophrenia201. Further molecular analyses of molecular patterns in psychiatric animal models need to be correlated with human studies to identify signatures associated with specific alterations on the genetic level or with the consequences of environmental insults. The use of biomarkers specific for different symptoms could help in new pharmacological approaches. For example current typical and atypical drugs fail mainly in targeting the negative and cognitive symptoms of schizophrenia11,182,202. The study of genetic and neurodevelopmental models could enable us to better identify risk biomarkers and help to initiate research programmes for discovery of novel treatments which target the fundamental pathogenesis instead of the current approach of symptom reduction. The relationship between genetic and environmental impact will probably remain the focus of current neuropsychiatric research over the next decade and the increased use of disease-relevant animal models could lead to new insights. The additional application of the RDoC system into modern approaches with psychiatric animal models could enhance the exchange between the preclinical and clinical stage and help to overcome the current stagnation in translational research.

Acknowledgements

This research was supported by the Stanley Medical Research Institute (SMRI), the European Union FP7 SchizDX research programme (grant reference 223427) and the NEWMEDS Innovative Medicines Initiative.

References

  • 1. WHO 2004. "The global burden of disease: 2004 update". Disponível em: <http://www.who.int/healthinfo/global_burden_disease/2004_report_update/en/index.html>
  • 2. McGrath JJ, Susser ES. New directions in the epidemiology of schizophrenia. Med J Aust. 2009;190(4 Suppl):S7-9.
  • 3. Nestler EJ, Hyman SE. Animal models of neuropsychiatric disorders. Nat Neurosci. 2010;13(10):1161-9.
  • 4. Goff DC, Coyle JT. The emerging role of glutamate in the pathophysiology and treatment of schizophrenia. Am J Psychiatry. 2001;158(9):1367-77.
  • 5. Jones P, Rodgers B, Murray R, Marmot M. Child development risk factors for adult schizophrenia in the British 1946 birth cohort. Lancet. 1994;344(8934):1398-402.
  • 6. Mueser KT, McGurk SR. Schizophrenia. Lancet. 2004;363(9426):2063-72.
  • 7. Rapoport JL, Addington AM, Frangou S, Psych MR. The neurodevelopmental model of schizophrenia: update 2005. Mol Psychiatry. 200510(5):434-49.
  • 8. Swerdlow NR, Braff DL, Taaid N, Geyer MA. Assessing the validity of an animal model of deficient sensorimotor gating in schizophrenic patients. Arch Gen Psychiatry. 1994;51(2):139-54.
  • 9. Marder SR, Fenton W. Measurement and Treatment Research to Improve Cognition in Schizophrenia: NIMH MATRICS initiative to support the development of agents for improving cognition in schizophrenia. Schizophr Res. 2004;72(1):5-9.
  • 10. Nuechterlein KH, Barch DM, Gold JM, Goldberg TE, Green MF, Heaton RK. Identification of separable cognitive factors in schizophrenia. Schizophr Res. 2004;72(1):29-39.
  • 11. Mintz J, Kopelowicz A. CUtLASS confirms CATIE. Arch Gen Psychiatry. 2007;64(8):978; author reply 979-980.
  • 12. Williams JH, Wellman NA, Geaney DP, Cowen PJ, Feldon J, Rawlins JN. Reduced latent inhibition in people with schizophrenia: an effect of psychosis or of its treatment. Br J Psychiatry. 1998;172:243-9.
  • 13. Rascle C, Mazas O, Vaiva G, Tournant M, Raybois O, Goudemand M, et al. Clinical features of latent inhibition in schizophrenia. Schizophr Res. 2001;51(2-3):149-61.
  • 14. Nieoullon A. Dopamine and the regulation of cognition and attention. Prog Neurobiol. 2002;67(1):53-83.
  • 15. Braff DL, Geyer MA. Sensorimotor gating and schizophrenia. Human and animal model studies. Arch Gen Psychiatry. 1990;47(2):181-8.
  • 16. Turetsky BI, Calkins ME, Light GA, Olincy A, Radant AD, Swerdlow NR. Neurophysiological endophenotypes of schizophrenia: the viability of selected candidate measures. Schizophr Bull. 2007;33(1):69-94.
  • 17. Swerdlow NR, Lipska BK, Weinberger DR, Braff DL, Jaskiw GE, Geyer MA. Increased sensitivity to the sensorimotor gating-disruptive effects of apomorphine after lesions of medial prefrontal cortex or ventral hippocampus in adult rats. Psychopharmacology (Berl). 1995;122(1):27-34.
  • 18. Colacicco G, Welzl H, Lipp HP, Würbel H. Attentional set-shifting in mice: modification of a rat paradigm, and evidence for strain-dependent variation. Behav Brain Res. 2002;132(1):95-102.
  • 19. McCarley RW, Wible CG, Frumin M, Hirayasu Y, Levitt JJ, Fischer IA, et al. MRI anatomy of schizophrenia. Biol Psychiatry. 1999;45(9):1099-119.
  • 20. Braff DL, Geyer MA, Swerdlow NR. Human studies of prepulse inhibition of startle: normal subjects, patient groups, and pharmacological studies. Psychopharmacology (Berl). 2001;156(2-3):234-58.
  • 21. Lewis DA, Hashimoto T, Volk DW. Cortical inhibitory neurons and schizophrenia. Nat Rev Neurosci. 2005;6(4):312-24.
  • 22. Insel T, Cuthbert B, Garvey M, Heinssen R, Pine DS, Quinn K, et al. Research domain criteria (RDoC): toward a new classification framework for research on mental disorders. Am J Psychiatry. 2010;167(7):748-51.
  • 23. Meltzer HY, Stahl SM. The dopamine hypothesis of schizophrenia: a review. Schizophr Bull. 1976;2(1):19-76.
  • 24. Coyle JT. Glutamate and schizophrenia: beyond the dopamine hypothesis. Cell Mol Neurobiol. 2006;26(4-6):365-84.
  • 25. Semkovska M, Bédard MA, Stip E. [Hypofrontality and negative symptoms in schizophrenia: synthesis of anatomic and neuropsychological knowledge and ecological perspectives]. Encephale. 2001;27(5):405-15.
  • 26. Boyer P, Phillips JL, Rousseau FL, Ilivitsky S. Hippocampal abnormalities and memory deficits: new evidence of a strong pathophysiological link in schizophrenia. Brain Res Rev. 2007;54(1):92-112.
  • 27. Geyer M, Markou A, editors. The role of preclinical models in the development of psychotropic drugs. Neuropsychopharmacology: The Fifth Generation of Progress. Washington, DC: American College of Neuropsychopharmacology; 2002.
  • 28. Van der Staay FJ, Arndt SS, Nordquist RE. Evaluation of animal models of neurobehavioral disorders. Behav Brain Funct. 2009;5:11.
  • 29. Kelly J, Murray RM. What risk factors tell us about the causes of schizophrenia and related psychoses. Curr Psychiatry Rep. 2000;2(5):378-85.
  • 30. Mäki P, Veijola J, Jones PB, Murray GK, Koponen H, Tienari P, et al. Predictors of schizophrenia: a review. Br Med Bull. 2005;73-74:1-15.
  • 31. Morgan C, Fisher H. Environment and schizophrenia: environmental factors in schizophrenia: childhood trauma: a critical review. Schizophr Bull. 2007;33(1):3-10.
  • 32. Kendler KS, Diehl SR. The genetics of schizophrenia: a current, genetic-epidemiologic perspective. Schizophr Bull. 1993;19(2):261-85.
  • 33. Sullivan PF, Kendler KS, Neale MC. Schizophrenia as a complex trait: evidence from a meta-analysis of twin studies. Arch Gen Psychiatry. 2003;60(12):1187-92.
  • 34. Maier W. Common risk genes for affective and schizophrenic psychoses. Eur Arch Psychiatry Clin Neurosci. 2008;258(Suppl 2):37-40.
  • 35. Allen NC, Bagade S, McQueen MB, Ioannidis JPA, Kavvoura FK, Khoury MJ, et al. Systematic meta-analyses and field synopsis of genetic association studies in schizophrenia: the SzGene database. Nat Genet. 2008;40(7):827-34.
  • 36. Costa E, Davis J, Grayson DR, Guidotti A, Pappas GD, Pesold C. Dendritic spine hypoplasticity and downregulation of reelin and GABAergic tone in schizophrenia vulnerability. Neurobiol Dis. 2001;8(5):723-42.
  • 37. O'Tuathaigh CM, Babovic D, O'Meara G, Clifford JJ, Croke DT, Waddington JL. Susceptibility genes for schizophrenia: characterisation of mutant mouse models at the level of phenotypic behaviour. Neurosci Biobehav Rev. 2007;31(1):60-78.
  • 38. Harrison PJ, Weinberger DR. Schizophrenia genes, gene expression, and neuropathology: on the matter of their convergence. Mol Psychiatry. 2005;10(1):40-68; image 45.
  • 39. Owen MJ, Craddock N, O'Donovan MC. Schizophrenia: genes at last? Trends Genet. 2005;21(9):518-25.
  • 40. Karayiorgou M, Gogos JA. Schizophrenia genetics: uncovering positional candidate genes. Eur J Hum Genet. 2006;14(5):512-9.
  • 41. St Clair D, Blackwood D, Muir W, Carothers A, Walker M, Spowart G, et al. Association within a family of a balanced autosomal translocation with major mental illness. Lancet. 1990;336(8706):13-6.
  • 42. Hwu HG, Liu CM, Fann CS, Ou-Yang WC, Lee SF. Linkage of schizophrenia with chromosome 1q loci in Taiwanese families. Mol Psychiatry. 2003;8(4):445-52.
  • 43. Jaaro-Peled H. Gene models of schizophrenia: DISC1 mouse models. Prog Brain Res. 2009;179:75-86.
  • 44. Clapcote SJ, Lipina TV, Millar JK, Mackie S, Christie S, Ogawa F, et al. Behavioral phenotypes of Disc1 missense mutations in mice. Neuron. 2007;54(3):387-402.
  • 45. Koike H, Arguello PA, Kvajo M, Karayiorgou M, Gogos JA. Disc1 is mutated in the 129S6/SvEv strain and modulates working memory in mice. Proc Natl Acad Sci U S A. 2006;103(10):3693-7.
  • 46. Konarski JZ, McIntyre RS, Grupp LA, Kennedy SH. Is the cerebellum relevant in the circuitry of neuropsychiatric disorders? J Psychiatry Neurosci. 2005;30(3):178-86.
  • 47. Ross CA, Margolis RL, Reading SA, Pletnikov M, Coyle JT. Neurobiology of schizophrenia. Neuron. 2006;52(1):139-53.
  • 48. Schurov IL, Handford EJ, Brandon NJ, Whiting PJ. Expression of disrupted in schizophrenia 1 (DISC1) protein in the adult and developing mouse brain indicates its role in neurodevelopment. Mol Psychiatry. 2004;9(12):1100-10.
  • 49. Arguello PA, Gogos JA. Modeling madness in mice: one piece at a time. Neuron. 2006;52(1):179-96.
  • 50. Terrin A, Di Benedetto G, Pertegato V, Cheung YF, Baillie G, Lynch MJ, et al. PGE(1) stimulation of HEK293 cells generates multiple contiguous domains with different [cAMP]: role of compartmentalized phosphodiesterases. J Cell Biol. 2006;175(3):441-51.
  • 51. Fatemi SH, King DP, Reutiman TJ, Folsom TD, Laurence JA, Lee S, et al. PDE4B polymorphisms and decreased PDE4B expression are associated with schizophrenia. Schizophr Res. 2008;101(1-3):36-49.
  • 52. Hikida T, Jaaro-Peled H, Seshadri S, Oishi K, Hookway C, Kong S, et al. Dominant-negative DISC1 transgenic mice display schizophrenia-associated phenotypes detected by measures translatable to humans. Proc Natl Acad Sci U S A. 2007;104(36):14501-16.
  • 53. Pletnikov MV, Ayhan Y, Nikolskaia O, Xu Y, Ovanesov MV, Huang H, et al. Inducible expression of mutant human DISC1 in mice is associated with brain and behavioral abnormalities reminiscent of schizophrenia. Mol Psychiatry. 2008;13(2):173-86, 115.
  • 54. Pogorelov VM, Nomura J, Kim J, Kannan G, Ayhan Y, Yang C, et al. Mutant DISC1 affects methamphetamine-induced sensitization and conditioned place preference: a comorbidity model. Neuropharmacology. 2012;62(3):1242-51. Epub 2011 Feb 17.
  • 55. Brandon NJ, Handford EJ, Schurov I, Rain JC, Pelling M, Duran-Jimeniz B, et al. Disrupted in Schizophrenia 1 and Nudel form a neurodevelopmentally regulated protein complex: implications for schizophrenia and other major neurological disorders. Mol Cell Neurosci. 2004;25(1):42-55.
  • 56. Eastwood SL. The synaptic pathology of schizophrenia: is aberrant neurodevelopment and plasticity to blame? Int Rev Neurobiol. 2004;59:47-72.
  • 57. Honer WG, Young CE. Presynaptic proteins and schizophrenia. Int Rev Neurobiol. 2004;59:175-99.
  • 58. Camargo LM, Collura V, Rain JC, Mizuguchi K, Hermjakob H, Kerrien S, et al. Disrupted in Schizophrenia 1 Interactome: evidence for the close connectivity of risk genes and a potential synaptic basis for schizophrenia. Mol Psychiatry. 2007;12(1):74-86.
  • 59. Arnold SE, Franz BR, Gur RC, Gur RE, Shapiro RM, Moberg P, et al. Smaller neuron size in schizophrenia in hippocampal subfields that mediate cortical-hippocampal interactions. Am J Psychiatry. 1995;152(5):738-48.
  • 60. Reif A, Fritzen S, Finger M, Strobel A, Lauer M, Schmitt A, et al. Neural stem cell proliferation is decreased in schizophrenia, but not in depression. Mol Psychiatry. 2006;11(5):514-22.
  • 61. Kvajo M, McKellar H, Arguelo PA, Drew LJ, Moore H, MacDermott AB, et al. A mutation in mouse Disc1 that models a schizophrenia risk allele leads to specific alterations in neuronal architecture and cognition. Proc Natl Acad Sci U S A. 2008;105(19):7076-81.
  • 62. Stefansson H, Sigurdsson E, Steinthorsdottir V, Bjornsdottir S, Sigmundsson T, Ghosh S, et al. Neuregulin 1 and susceptibility to schizophrenia. Am J Hum Genet. 2002;71(4):877-92.
  • 63. Harrison PJ, Law AJ. Neuregulin 1 and schizophrenia: genetics, gene expression, and neurobiology. Biol Psychiatry. 2006;60(2):32-40.
  • 64. Mei L, Xiong WC. Neuregulin 1 in neural development, synaptic plasticity and schizophrenia. Nat Rev Neurosci. 2008;9(6):437-52.
  • 65. Van den Buuse M, Wischhof L, Lee RX, Martin S, Karl T. Neuregulin 1 hypomorphic mutant mice: enhanced baseline locomotor activity but normal psychotropic drug-induced hyperlocomotion and prepulse inhibition regulation. Int J Neuropsychopharmacol. 2009;12(10):1383-93.
  • 66. Duffy L, Cappas E, Scimone A, Schofield PR, Karl T. Behavioral profile of a heterozygous mutant mouse model for EGF-like domain neuregulin 1. Behav Neurosci. 2008;122(4):748-59.
  • 67. Deakin IH, Law AJ, Oliver PL, Schwab MH, Nave KA, Harrison PJ, et al. Behavioural characterization of neuregulin 1 type I overexpressing transgenic mice. Neuroreport. 2009;20(17):1523-8.
  • 68. Wang XD, Su YA, Guo CM, Yang Y, Si TM. Chronic antipsychotic drug administration alters the expression of neuregulin 1beta, ErbB2, ErbB3, and ErbB4 in the rat prefrontal cortex and hippocampus. Int J Neuropsychopharmacol. 2008;11(4):553-61.
  • 69. Rimer M, Barrett DW, Maldonado MA, Vock VM, Gonzalez-Lima F. Neuregulin-1 immunoglobulin-like domain mutant mice: clozapine sensitivity and impaired latent inhibition. Neuroreport. 2005;16(3):271-5.
  • 70. O'Tuathaigh CM, Babovic D, O'Sullivan GJ, Clifford JJ, Tighe O, et al. Phenotypic characterization of spatial cognition and social behavior in mice with 'knockout' of the schizophrenia risk gene neuregulin 1. Neuroscience. 2007;147(1):18-27.
  • 71. O'Tuathaigh CM, O'Connor AM, O'Sullivan GJ, Lai D, Harvey R, Croke DT, et al. Disruption to social dyadic interactions but not emotional/anxiety-related behaviour in mice with heterozygous 'knockout' of the schizophrenia risk gene neuregulin-1. Prog Neuropsychopharmacol Biol Psychiatry. 2008;32(2):462-6.
  • 72. Ehrlichman RS, Luminais SN, White SL, Rudnick ND, Ma N, Dow HC, et al. Neuregulin 1 transgenic mice display reduced mismatch negativity, contextual fear conditioning and social interactions. Brain Res. 2009;1294:116-27.
  • 73. Bjarnadottir M, Misner DL, Haverfield-Gross S, Bruun S, Helgason VG, Stefansson H, et al. Neuregulin1 (NRG1) signaling through Fyn modulates NMDA receptor phosphorylation: differential synaptic function in NRG1+/- knock-outs compared with wild-type mice. J Neurosci. 2007;27(17):4519-29.
  • 74. Green MF, Nuechterlein KH, Gold JM, Barch DM, Cohen J, Essock S, et al. Approaching a consensus cognitive battery for clinical trials in schizophrenia: the NIMH-MATRICS conference to select cognitive domains and test criteria. Biol Psychiatry. 2004;56(5):301-7.
  • 75. Paterlini M, Zakharenko SS, Lai WS, Qin J, Zhang H, Mukai J, et al. Transcriptional and behavioral interaction between 22q11.2 orthologs modulates schizophrenia-related phenotypes in mice. Nat Neurosci. 2005;8(11):1586-94.
  • 76. Sobin C, Kiley-Brabeck K, Daniels S, Khuri J, Taylor L, Blundell M, et al. Neuropsychological characteristics of children with the 22q11 Deletion Syndrome: a descriptive analysis. Child Neuropsychol. 2005;11(1):39-53.
  • 77. Pulver AE, Nestadt G, Goldberg R, Shprintzen RJ, Lamacz M, Wolyniec PS, et al. Psychotic illness in patients diagnosed with velo-cardio-facial syndrome and their relatives. J Nerv Ment Dis. 1994;182(8):476-8.
  • 78. Murphy KC, Jones LA, Owen MJ. High rates of schizophrenia in adults with velo-cardio-facial syndrome. Arch Gen Psychiatry. 1999;56(10):940-45.
  • 79. Bassett AS, Chow EW, Husted J, Weksberg R, Caluseriu O, Webb GD, et al. Clinical features of 78 adults with 22q11 Deletion Syndrome. Am J Med Genet A. 2005;138(4):307-13.
  • 80. Karayiorgou M, Morris MA, Morrow B, Shprintzen RJ, Goldberg C, Borrow J, et al. Schizophrenia susceptibility associated with interstitial deletions of chromosome 22q11. Proc Natl Acad Sci U S A. 1995;92(17):7612-6.
  • 81. Usiskin SI, Nicolson R, Krasnewich DM, Yan W, Lenane M, Wudarsky M, et al. Velocardiofacial syndrome in childhood-onset schizophrenia. J Am Acad Child Adolesc Psychiatry. 1999;38(12):1536-43.
  • 82. Pulver AE, Karayiorgou M, Lasseter VK, Wolyniec P, Kasch L, Antonarakis S, et al. Follow-up of a report of a potential linkage for schizophrenia on chromosome 22q12-q13.1: Part 2. Am J Med Genet. 1994;54(1):44-50.
  • 83. Pulver AE, Karayiorgou M, Wolyniec PS, Lasseter VK, Kasch L, Nestadt G, et al. Sequential strategy to identify a susceptibility gene for schizophrenia: report of potential linkage on chromosome 22q12-q13.1: Part 1. Am J Med Genet. 1994;54(1):36-43.
  • 84. Lasseter VK, Pulver AE, Wolyniec PS, Nestadt G, Meyers D, Karayiorgou M, et al. Follow-up report of potential linkage for schizophrenia on chromosome 22q: Part 3. Am J Med Genet. 1995;60(2):172-3.
  • 85. Gogos JA, Santha M, Takacs Z, Beck KD, Luine V, Lucas LR, et al. The gene encoding proline dehydrogenase modulates sensorimotor gating in mice. Nat Genet. 1999;21(4):434-9.
  • 86. Fremeau RT Jr, Caron MG, Blakely RD. Molecular cloning and expression of a high affinity L-proline transporter expressed in putative glutamatergic pathways of rat brain. Neuron. 1992;8(5):915-26.
  • 87. Cohen SM, JV Nadler. Proline-induced potentiation of glutamate transmission. Brain Res. 1997;761(2):271-82.
  • 88. Renick SE, Kleven DT, Chan J, Stenius K, Milner TA, Pickel VM, et al. The mammalian brain high-affinity L-proline transporter is enriched preferentially in synaptic vesicles in a subpopulation of excitatory nerve terminals in rat forebrain. J Neurosci. 1999;19(1):21-33.
  • 89. Kimber WL, Hsieh P, Hirotsune S, Yuva-Paylor L, Sutherland HF, Chen A, et al. Deletion of 150 kb in the minimal DiGeorge/velocardiofacial syndrome critical region in mouse. Hum Mol Genet. 1999;8(12):2229-37.
  • 90. Paylor R, McIlwain KL, McAninch R, Nellis A, Yuva-Paylor LA, Baldini A, et al. Mice deleted for the DiGeorge/velocardiofacial syndrome region show abnormal sensorimotor gating and learning and memory impairments. Hum Mol Genet. 2001;10(23):2645-50.
  • 91. Long JM, LaPorte P, Merscher S, Funke B, Saint-Jore B, Puech A, et al. Behavior of mice with mutations in the conserved region deleted in velocardiofacial/DiGeorge syndrome. Neurogenetics. 2006;7(4):247-57.
  • 92. Mukai J, Dhilla A, Drew LJ, Stark KL, Cao L, MacDermott AB, et al. Palmitoylation-dependent neurodevelopmental deficits in a mouse model of 22q11 microdeletion. Nat Neurosci. 2008;11(11):1302-10.
  • 93. Stark KL, Xu B, Bagchi A, Lai WS, Liu H, Hsu R, et al. Altered brain microRNA biogenesis contributes to phenotypic deficits in a 22q11-deletion mouse model. Nat Genet. 2008;40(6):751-60.
  • 94. Meechan DW, Tucker ES, Maynard TM, LaMantia AS. Diminished dosage of 22q11 genes disrupts neurogenesis and cortical development in a mouse model of 22q11 deletion/DiGeorge syndrome. Proc Natl Acad Sci U S A. 2009;106(38):16434-45.
  • 95. Karayiorgou M, Simon TJ, Gogos JA. 22q11.2 microdeletions: linking DNA structural variation to brain dysfunction and schizophrenia. Nat Rev Neurosci. 2010;11(6):402-16.
  • 96. Giros B, Jaber M, Jones SR, Wightman RM, Caron MG. Hyperlocomotion and indifference to cocaine and amphetamine in mice lacking the dopamine transporter. Nature. 1996;379(6566):606-12.
  • 97. Gainetdinov RR, Jones SR, Fumagalli F, Wightman RM, Caron MG. Re-evaluation of the role of the dopamine transporter in dopamine system homeostasis. Brain Res Brain Res Rev. 1998;26(2-3):148-53.
  • 98. Dzirasa K, Santos LM, Ribeiro R, Stapleton J, Gainetdinov RR, Caron MG, et al. Persistent hyperdopaminergia decreases the peak frequency of hippocampal theta oscillations during quiet waking and REM sleep. PLoS One. 2009;4(4):e5238.
  • 99. Hall FS, Li XF, Randall-Thompson J, Sora I, Murphy DL, Lesch KP, et al. Cocaine-conditioned locomotion in dopamine transporter, norepinephrine transporter and 5-HT transporter knockout mice. Neuroscience. 2009;162(4):870-80.
  • 100. Pogorelov VM, Rodriguiz RM, Insco ML, Caron MG, Wetsel WC. Rodriguiz, et al. Novelty seeking and stereotypic activation of behavior in mice with disruption of the Dat1 gene. Neuropsychopharmacology. 2005;30(10):1818-31.
  • 101. Wu N, Cepeda C, Zhuang X, Levine MS. Altered corticostriatal neurotransmission and modulation in dopamine transporter knock-down mice. J Neurophysiol. 2007;98(1):423-32. Epub 2007 May 23.
  • 102. Bossé R, Fumagalli F, Jaber M, Giros B, Gainetdinov RR, Wetsel WC, et al. Anterior pituitary hypoplasia and dwarfism in mice lacking the dopamine transporter. Neuron. 1997;19(1):127-38.
  • 103. Upadhyaya AR, El-Sheikh R, MacMaster FP, Diwadkar VA, Keshavan MS. Pituitary volume in neuroleptic-naive schizophrenia: a structural MRI study. Schizophr Res. 2007;90(1-3):266-73.
  • 104. Mohn AR, Gainetdinov RR, Caron MG, Koller BH. Mice with reduced NMDA receptor expression display behaviors related to schizophrenia. Cell. 1999;98(4):427-36.
  • 105. Moy SS, Perez A, Koller BH, Duncan GE. Amphetamine-induced disruption of prepulse inhibition in mice with reduced NMDA receptor function. Brain Res. 2006;1089(1):186-94.
  • 106. Bickel S, Lipp HP, Umbricht D. Early auditory sensory processing deficits in mouse mutants with reduced NMDA receptor function. Neuropsychopharmacology. 2008;33(7):1680-9.
  • 107. Korotkova T, Fuchs EC, Ponomarenko A, Von Engelhardt J, Monyer H. NMDA receptor ablation on parvalbumin-positive interneurons impairs hippocampal synchrony, spatial representations, and working memory. Neuron. 2010;68(3):557-69.
  • 108. Geyer MA, Swerdlow NR, Mansbach RS, Braff D. Startle response models of sensorimotor gating and habituation deficits in schizophrenia. Brain Res Bull. 1990;25(3):485-98.
  • 109. Lodge DJ, Behrens MM, Grace AA. A loss of parvalbumin-containing interneurons is associated with diminished oscillatory activity in an animal model of schizophrenia. J Neurosci. 2009;29(8):2344-54.
  • 110. Belforte JE, Zsiros V, Sklar ER, Jiang Z, Yu G, Li Y, et al. Postnatal NMDA receptor ablation in corticolimbic interneurons confers schizophrenia-like phenotypes. Nat Neurosci. 2010;13(1):76-83.
  • 111. Tsuang M. Schizophrenia: genes and environment. Biol Psychiatry. 2000;47(3):210-20.
  • 112. Harrison PJ. The neuropathology of schizophrenia. A critical review of the data and their interpretation. Brain. 1999;122(Pt 4):593-624.
  • 113. Lipska BK, Jaskiw GE, Weinberger DR. Postpubertal emergence of hyperresponsiveness to stress and to amphetamine after neonatal excitotoxic hippocampal damage: a potential animal model of schizophrenia. Neuropsychopharmacology. 1993.9(1):67-75.
  • 114. Moore H, Jentsch JD, Ghajarnia M, Geyer MA, Grace AA. A neurobehavioral systems analysis of adult rats exposed to methylazoxymethanol acetate on E17: implications for the neuropathology of schizophrenia. Biol Psychiatry. 2006;60(3):253-64.
  • 115. Berger N, Vaillancourt C, Boksa P. Genetic factors modulate effects of C-section birth on dopaminergic function in the rat. Neuroreport. 2000;11(3):639-43.
  • 116. Fone KC, Porkess MV. Behavioural and neurochemical effects of post-weaning social isolation in rodents-relevance to developmental neuropsychiatric disorders. Neurosci Biobehav Rev. 2008;32(6):1087-102.
  • 117. Markham JA, Koenig JI. Prenatal stress: role in psychotic and depressive diseases. Psychopharmacology (Berl). 2011;214(1):89-106.
  • 118. Fatemi SH, Emamian ES, Kist D, Sidwell RW, Nakajima K, Akhter P, et al. Defective corticogenesis and reduction in Reelin immunoreactivity in cortex and hippocampus of prenatally infected neonatal mice. Mol Psychiatry. 1999;4(2):145-54.
  • 119. Steiger JL, Alexander MJ, Galler JR, Farb DH, Russek SJ. Effects of prenatal malnutrition on GABAA receptor alpha1, alpha3 and beta2 mRNA levels. Neuroreport. 2003;14(13):1731-5.
  • 120. Flagstad P, Mørk A, Glenthøj BY, Van Beek J, Michael-Titus AT, Didriksen M. Disruption of neurogenesis on gestational day 17 in the rat causes behavioral changes relevant to positive and negative schizophrenia symptoms and alters amphetamine-induced dopamine release in nucleus accumbens. Neuropsychopharmacology. 2004;29(11):2052-64.
  • 121. Palmer AA, Printz DJ, Butler PD, Dulawa SC, Printz MP. Prenatal protein deprivation in rats induces changes in prepulse inhibition and NMDA receptor binding. Brain Res. 2004;996(2):193-201.
  • 122. Flagstad P, Glenthøj BY, Didriksen M. Cognitive deficits caused by late gestational disruption of neurogenesis in rats: a preclinical model of schizophrenia. Neuropsychopharmacology. 2005;30(2):250-60.
  • 123. Meyer U, Nyffeler M, Schwendener S, Knuesel I, Yee BK, Feldon J. Relative prenatal and postnatal maternal contributions to schizophrenia-related neurochemical dysfunction after in utero immune challenge. Neuropsychopharmacology. 2008;33(2):441-56.
  • 124. Tseng KY, Lewis BL, Hashimoto T, Sesack SR, Kloc M, Lewis DA, et al. A neonatal ventral hippocampal lesion causes functional deficits in adult prefrontal cortical interneurons. J Neurosci. 2008;28(48):12691-9.
  • 125. Mednick SA, Machon RA, Huttunen MO, Bonett D. Adult schizophrenia following prenatal exposure to an influenza epidemic. Arch Gen Psychiatry. 1988;45(2):189-92.
  • 126. Mortensen PB, Pedersen CB, Westergaard T, Wohlfahrt J, Ewald H, Mors O, et al. Effects of family history and place and season of birth on the risk of schizophrenia. N Engl J Med. 1999;340(8):603-8.
  • 127. Davies G, Welham J, Chant D, Torrey EF, Torrey J. A systematic review and meta-analysis of Northern Hemisphere season of birth studies in schizophrenia. Schizophr Bull. 2003;29(3):587-93.
  • 128. Romero E, Guaza C, Castellano B, Borrell J. Ontogeny of sensorimotor gating and immune impairment induced by prenatal immune challenge in rats: implications for the etiopathology of schizophrenia. Mol Psychiatry. 2010;15(4):372-83.
  • 129. Kozlovsky N, Belmaker RH, Agam G. Low GSK-3beta immunoreactivity in post-mortem frontal cortex of schizophrenic patients. Am J Psychiatry. 2000;157(5):831-3.
  • 130. Kozlovsky N, Belmaker RH, Agam G. Low GSK-3 activity in frontal cortex of schizophrenic patients. Schizophr Res. 2001;52(1-2):101-5.
  • 131. Kozlovsky N, Shanon-Weickert C, Tomaskovic-Crook E, Kleinman JE, Belmaker RH, Agam G. Reduced GSK-3beta mRNA levels in post-mortem dorsolateral prefrontal cortex of schizophrenic patients. J Neural Transm. 2004;111(12):1583-92.
  • 132. Monteleone P, Fabrazzo M, Tortorella A, Maj M. Plasma levels of interleukin-6 and tumor necrosis factor alpha in chronic schizophrenia: effects of clozapine treatment. Psychiatry Res. 1997;71(1):11-7.
  • 133. Gaughran F. Immunity and schizophrenia: autoimmunity, cytokines, and immune responses. Int Rev Neurobiol. 2002;52:275-302.
  • 134. Skurkovich SV, Aleksandrovsky YA, Chekhonin VP, Ryabukhin IA, Chakhava KO, Skurkovich B. Improvement in negative symptoms of schizophrenia with antibodies to tumor necrosis factor-alpha and to interferon-gamma: a case report. J Clin Psychiatry. 2003;64(6):734-5.
  • 135. Romero E, Ali C, Molina-Holgado E, Castellano B, Guaza C, Borrell J. Neurobehavioral and immunological consequences of prenatal immune activation in rats. Influence of antipsychotics. Neuropsychopharmacology. 2007;32(8):1791-804.
  • 136. Shi L, Fatemi SH, Sidwell RW, Patterson PH. Maternal influenza infection causes marked behavioral and pharmacological changes in the offspring. J Neurosci. 2003;23(1):297-302.
  • 137. Cotter D, Takei N, Farrell M, Sham P, Quinn P, Larkin C, et al. Does prenatal exposure to influenza in mice induce pyramidal cell disarray in the dorsal hippocampus? Schizophr Res. 1995;16(3):233-41.
  • 138. Altshuler LL, Conrad A, Kovelman JA, Scheibel A. Hippocampal pyramidal cell orientation in schizophrenia. A controlled neurohistologic study of the Yakovlev collection. Arch Gen Psychiatry. 1987;44(12):1094-8.
  • 139. Christison GW, Casanova MF, Weinberger DR, Rawlings R, Kleinman JE. A quantitative investigation of hippocampal pyramidal cell size, shape, and variability of orientation in schizophrenia. Arch Gen Psychiatry. 1989;46(11):1027-32.
  • 140. Benes FM, Sorensen I, Bird ED. Reduced neuronal size in posterior hippocampus of schizophrenic patients. Schizophr Bull. 1991;17(4):597-608.
  • 141. Wright P, Murray RM. Schizophrenia: prenatal influenza and autoimmunity. Ann Med. 1993;25(5):497-502.
  • 142. Müller N, Riedel M, Ackenheil M, Schwarz MJ. The role of immune function in schizophrenia: an overview. Eur Arch Psychiatry Clin Neurosci. 1999;249(Suppl 4):62-8.
  • 143. Nawa H, Takahashi M, Patterson PH. Cytokine and growth factor involvement in schizophrenia: support for the developmental model. Mol Psychiatry. 2000;5(6):594-603.
  • 144. Boin F, Zanardini R, et al. Association between -G308A tumor necrosis factor alpha gene polymorphism and schizophrenia. Mol Psychiatry. 2001;6(1):79-82.
  • 145. Rothermundt M, Arolt V, Bayer TA. Review of immunological and immunopathological findings in schizophrenia. Brain Behav Immun. 2001;15(4):319-339.
  • 146. Susser ES, Lin SP. Schizophrenia after prenatal exposure to the Dutch Hunger Winter of 1944-1945. Arch Gen Psychiatry. 1992;49(12):983-8.
  • 147. Susser E, Neugebauer R, Hoek HW, Brown AS, Lin S, Labovitz D, et al. Schizophrenia after prenatal famine. Further evidence. Arch Gen Psychiatry. 1996;53(1):25-31.
  • 148. St Clair D, Xu M, Wang P, Yu Y, Fang Y, Zhang F, et al. Rates of adult schizophrenia following prenatal exposure to the Chinese famine of 1959-1961. JAMA. 2005;294(5):557-62.
  • 149. Palmer AA, Brown AS, Keegan D, Siska LD, Susser E, Rotrosen J, et al. Prenatal protein deprivation alters dopamine-mediated behaviors and dopaminergic and glutamatergic receptor binding. Brain Res. 2008;1237:62-74.
  • 150. Chen JC, Turiak G, Galler J, Volicer L. Postnatal changes of brain monoamine levels in prenatally malnourished and control rats. Int J Dev Neurosci. 1997;15(2):257-63.
  • 151. Stone JM, Morrison PD, Pilowsky LS. Glutamate and dopamine dysregulation in schizophrenia: a synthesis and selective review. J Psychopharmacol. 2007;21(4):440-52.
  • 152. Tonkiss J, Galler JR. Prenatal protein malnutrition and working memory performance in adult rats. Behav Brain Res. 1990;40(2):95-107.
  • 153. Barrett DE, Radke-Yarrow M. Effects of nutritional supplementation on children's responses to novel, frustrating, and competitive situations. Am J Clin Nutr. 1985;42(1):102-20.
  • 154. Almeida SS, Tonkiss J, Galler JR. Prenatal protein malnutrition affects the social interactions of juvenile rats. Physiol Behav. 1996;60(1):197-201.
  • 155. Marichich ES, Molina VA, Orsingher OA. Persistent changes in central catecholaminergic system after recovery of perinatally undernourished rats. J Nutr. 1979;109(6):1045-50.
  • 156. Bayer S, Altman J, editors. Neurogenesis and neuronal migration. In: Paxinos G, editor. The rat nervous system. 2nd San Diego, CA: Academic Press; 1995. p. 1041-78.
  • 157. Gourevitch R, Rocher C, Le Pen G, Krebs MO, Jay TM. Working memory deficits in adult rats after prenatal disruption of neurogenesis. Behav Pharmacol. 2004;15(4):287-92.
  • 158. Fiore M, Talamini L, Angelucci F, Koch T, Aloe L, Korf J. Prenatal methylazoxymethanol acetate alters behavior and brain NGF levels in young rats: a possible correlation with the development of schizophrenia-like deficits. Neuropharmacology. 1999;38(6):857-69.
  • 159. Tremolizzo L, Carboni G, Ruzicka WB, Mitchell CP, Sugaya I, Tusting P, et al. An epigenetic mouse model for molecular and behavioral neuropathologies related to schizophrenia vulnerability. Proc Natl Acad Sci U S A. 2002;99(26):17095-100.
  • 160. Dong E, Agis-Balboa RC, Simonini MV, Grayson DR, Costa E, Guidotti A. Reelin and glutamic acid decarboxylase67 promoter remodeling in an epigenetic methionine-induced mouse model of schizophrenia. Proc Natl Acad Sci U S A. 2005;102(35):12578-83.
  • 161. Selemon LD, Rajkowska G, Goldman-Rakic PS. Abnormally high neuronal density in the schizophrenic cortex. A morphometric analysis of prefrontal area 9 and occipital area 17. Arch Gen Psychiatry. 1995;52(10):805-18; discussion 819-20.
  • 162. Thune JJ, Pakkenberg B. Stereological studies of the schizophrenic brain. Brain Res Brain Res Rev. 2000;31(2-3):200-4.
  • 163. Wright IC, Rabe-Hesketh S, Woodruff PW, David AS, Murray RM, Bullmore ET. Meta-analysis of regional brain volumes in schizophrenia. Am J Psychiatry. 2000;157(1):16-25.
  • 164. Ananth H, Popescu I, Critchley HD, Good CD, Frackowiak RS, Dolan RJ. Cortical and subcortical gray matter abnormalities in schizophrenia determined through structural magnetic resonance imaging with optimized volumetric voxel-based morphometry. Am J Psychiatry. 2002;159(9):1497-505.
  • 165. Selemon LD, Mrzljak J, Kleinman JE, Herman MM, Goldman-Rakic PS. Regional specificity in the neuropathologic substrates of schizophrenia: a morphometric analysis of Broca's area 44 and area 9. Arch Gen Psychiatry. 2003;60(1):69-77.
  • 166. Jönsson SA, Luts A, Guldberg-Kjaer N, Ohman R. Pyramidal neuron size in the hippocampus of schizophrenics correlates with total cell count and degree of cell disarray. Eur Arch Psychiatry Clin Neurosci. 1999;249(4):169-73.
  • 167. Hradetzky E, Sanderson TM, Tsang TM, Sherwood JL, Fitzjohn SM, Lakics V, et al. The methylazoxymethanol acetate (MAM-E17) rat model: molecular and functional effects in the hippocampus. Neuropsychopharmacology. 2012;37(2):364-77.
  • 168. Braff DL, Grillon C, Geyer MA. Gating and habituation of the startle reflex in schizophrenic patients. Arch Gen Psychiatry. 1992;49(3):206-15.
  • 169. Le Pen G, Gourevitch R, Hazane F, Hoareau C, Jay TM, Krebs MO. Peri-pubertal maturation after developmental disturbance: a model for psychosis onset in the rat. Neuroscience. 2006;143(2):395-405.
  • 170. Aleman A, Hijman R, De Haan EH, Kahn RS. Memory impairment in schizophrenia: a meta-analysis. Am J Psychiatry. 1999;156(9):1358-66.
  • 171. Heckers S, Curran T, Goff D, Rauch SL, Fischman AJ, Alpert NM, et al. Abnormalities in the thalamus and prefrontal cortex during episodic object recognition in schizophrenia. Biol Psychiatry. 2000;48(7):651-7.
  • 172. Gray JA, Joseph MH, Hemsley DR, Young AM, Warburton EC, Boulenguez P, et al. The role of mesolimbic dopaminergic and retrohippocampal afferents to the nucleus accumbens in latent inhibition: implications for schizophrenia. Behav Brain Res. 1995;71(1-2):19-31.
  • 173. Gray JA, Moran PM, Grigoryan G, Peters SL, Young AM, Joseph MH. Latent inhibition: the nucleus accumbens connection revisited. Behav Brain Res. 1997;88(1):27-34.
  • 174. Stark H, Rothe T, Wagner T, Scheich H. Learning a new behavioral strategy in the shuttle-box increases prefrontal dopamine. Neuroscience. 2004;126(1):21-9.
  • 175. Lavin A, Grace A. Effects of afferent stimulation and DA application on prefrontal cortical cells recorded intracellularly in vivo: comparisons between intact rats and rats with pharmacologically-induced disruption of cortical development. Soc Neurosci Abstr. 1997;23:2080.
  • 176. Kosmidis MH, Breier A, Fantie BD. Avoidance learning in schizophrenia: a dissociation between the effects of aversive and non-aversive stimuli. Schizophr Res. 1999;38(1):51-9.
  • 177. Le Pen G, Moreau JL. Disruption of prepulse inhibition of startle reflex in a neurodevelopmental model of schizophrenia: reversal by clozapine, olanzapine and risperidone but not by haloperidol. Neuropsychopharmacology. 2002;27(1):1-11.
  • 178. Lipska B, Weinberger D, editors. Cortical regulation of the mesolimbic dopamine system implications for schizophrenia. In: Kalivas PW, Barnes CD, editors. Limbic circuits and neuropsychiatry. Boca Raton: CRC Press; 1993.
  • 179. Lipska BK, Weinberger DR. Delayed effects of neonatal hippocampal damage on haloperidol-induced catalepsy and apomorphine-induced stereotypic behaviors in the rat. Brain Res Dev Brain Res. 1993;75(2):213-22.
  • 180. Chambers RA, Moore J, McEvoy JP, Levin ED. Cognitive effects of neonatal hippocampal lesions in a rat model of schizophrenia. Neuropsychopharmacology. 1996;15(6):587-94.
  • 181. Sams-Dodd F, Lipska BK, Weinberger DR. Neonatal lesions of the rat ventral hippocampus result in hyperlocomotion and deficits in social behaviour in adulthood. Psychopharmacology (Berl). 1997;132(3):303-10.
  • 182. Nuechterlein KH. Childhood precursors of adult schizophrenia. J Child Psychol Psychiatry. 1986;27(2):133-44.
  • 183. Auerbach JG, Hans S, Marcus J. Neurobehavioral functioning and social behavior of children at risk for schizophrenia. Isr J Psychiatry Relat Sci. 1993;30(1):40-9.
  • 184. Done DJ, Crow TJ, Johnstone EC, Sacker A. Childhood antecedents of schizophrenia and affective illness: social adjustment at ages 7 and 11. BMJ. 1994;309(6956):699-703.
  • 185. 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(2):145-60.
  • 186. Stine CD, Lu W, Wolf ME. Expression of AMPA receptor flip and flop mRNAs in the nucleus accumbens and prefrontal cortex after neonatal ventral hippocampal lesions. Neuropsychopharmacology. 2001;24(3):253-66.
  • 187. O'Donnell P, Lewis BL, Weinberger DR, Lipska BK. Neonatal hippocampal damage alters electrophysiological properties of prefrontal cortical neurons in adult rats. Cereb Cortex. 2002;12(9):975-82.
  • 188. Hashimoto T, Volk DW, Eggan SM, Mirnics K, Pierri JN, Sun Z, et al. Gene expression deficits in a subclass of GABA neurons in the prefrontal cortex of subjects with schizophrenia. J Neurosci. 2003;23(15):6315-26.
  • 189. Tordjman S, Drapier D, Bonnot O, Graignic R, Fortes S, Cohen D. Animal models relevant to schizophrenia and autism: validity and limitations. Behav Genet. 2007;37(1):61-78.
  • 190. Kluge W, Alsaif M, Guest PC, Schwarz E, Bahn S. Translating potential biomarker candidates for schizophrenia and depression to animal models of psychiatric disorders. Expert Rev Mol Diagn. 2011;11(7):721-33.
  • 191. Pechnick RN, George R, Poland RE. Characterization of the effects of the acute and repeated administration of MK-801 on the release of adrenocorticotropin, corticosterone and prolactin in the rat. Eur J Pharmacol. 1989;164(2):257-63.
  • 192. Issa G, Wilson C, Terry AV Jr, Pillai A. An inverse relationship between cortisol and BDNF levels in schizophrenia: data from human post-mortem and animal studies. Neurobiol Dis. 2010;39(3):327-33.
  • 193. Borrell J, Vela JM, Arévalo-Martin A, Molina-Holgado E, Guaza C. Prenatal immune challenge disrupts sensorimotor gating in adult rats. Implications for the etiopathogenesis of schizophrenia. Neuropsychopharmacology. 2002;26(2):204-15.
  • 194. Mitsuhashi S, Fukushima T, Tomiya M, Santa T, Imai K, Toyo'oka T. Determination of kynurenine levels in rat plasma by high-performance liquid chromatography with pre-column fluorescence derivatization. Anal Chim Acta. 2007;584(2):315-21.
  • 195. Tomiya M, Fukushima T, Kawai J, Aoyama C, Mitsuhashi S, Santa T, et al. Alterations of plasma and cerebrospinal fluid glutamate levels in rats treated with the N-methyl-D-aspartate receptor antagonist, ketamine. Biomed Chromatogr. 2006;20(6-7):628-33.
  • 196. Minassian A, Henry BL, Geyer MA, Paulus MP, Young JW, Perry W. The quantitative assessment of motor activity in mania and schizophrenia. J Affect Disord. 2010;120(1-3):200-6.
  • 197. Hellhammer D, Hero T, Gerhards F, Hellhammer J. Neuropattern: a new translational tool to detect and treat stress pathology. I. Strategical consideration. Stress. 2012;15(5):479-87. doi: 10.3109/10253890.2011.644604. Epub 2012 Jan 10.
  • 198. Hennekam RC. Hereditary multiple exostoses. J Med Genet. 1991;28(4):262-6.
  • 199. Sarnyai Z, Alsaif M, Bahn S, Ernst A, Guest PC, Hradetzky E, et al. Behavioral and molecular biomarkers in translational animal models for neuropsychiatric disorders. Int Rev Neurobiol. 2011;101:203-38.
  • 200. Levin Y, Wang L, Schwarz E, Koethe D, Leweke FM, Bahn S. Global proteomic profiling reveals altered proteomic signature in schizophrenia serum. Mol Psychiatry. 2010;15(11):1088-100.
  • 201. Schwarz E, Izmailov R, Spain M, Barnes A, Mapes JP, Guest PC, et al. Validation of a blood-based laboratory test to aid in the confirmation of a diagnosis of schizophrenia. Biomark Insights. 2010;5:39-47.
  • 202. Keefe RS, Bilder RM, Davis SM, Harvey PD, Palmer BW, Gold JM, et al. Neurocognitive effects of antipsychotic medications in patients with chronic schizophrenia in the CATIE Trial. Arch Gen Psychiatry. 2007;64(6):633-47.
  • Endereço para correspondência:

    Sabine Bahn
    University of Cambridge
    Tennis Court Road, Cambridge. CB2 1QT, UK
    Telephone: +44 1223 334 151 Fax: +44 1223 334 162
    Email:
  • Publication Dates

    • Publication in this collection
      14 Dec 2012
    • Date of issue
      2013

    History

    • Received
      23 Sept 2012
    • Accepted
      07 Nov 2012
    Faculdade de Medicina da Universidade de São Paulo Rua Ovídio Pires de Campos, 785 , 05403-010 São Paulo SP Brasil, Tel./Fax: +55 11 2661-8011 - São Paulo - SP - Brazil
    E-mail: archives@usp.br