Skip to main content
Journal of Psychiatry & Neuroscience : JPN logoLink to Journal of Psychiatry & Neuroscience : JPN
. 2011 Jul;36(4):E5–E29. doi: 10.1503/jpn.100054

Contribution of nonprimate animal models in understanding the etiology of schizophrenia

Noah L Lazar 1,, Richard WJ Neufeld 1, Donald P Cain 1
PMCID: PMC3120891  PMID: 21247514

Abstract

Schizophrenia is a severe psychiatric disorder that is characterized by positive and negative symptoms and cognitive impairments. The etiology of the disorder is complex, and it is thought to follow a multifactorial threshold model of inheritance with genetic and neurodevelopmental contributions to risk. Human studies are particularly useful in capturing the richness of the phenotype, but they are often limited to the use of correlational approaches. By assessing behavioural abnormalities in both humans and rodents, nonprimate animal models of schizophrenia provide unique insight into the etiology and mechanisms of the disorder. This review discusses the phenomenology and etiology of schizophrenia and the contribution of current nonprimate animal models with an emphasis on how research with models of neurotransmitter dysregulation, environmental risk factors, neurodevelopmental disruption and genetic risk factors can complement the literature on schizophrenia in humans.

Introduction

Schizophrenia is a devastating psychiatric disorder that disrupts cognition, emotion, language and thought, and it typically affects 0.5%–1.5% of the population.1 The onset of the disorder usually occurs during adolescence or early adulthood, although a prodromal phase usually precedes its onset.1 The etiology of the disorder is complex and is thought to be largely genetic in nature. Association and linkage studies with humans, although useful in determining genes that increase the risk for schizophrenia, have produced inconsistent results. Animal models may be particularly useful for assessing the impact of genetic and environmental risk factors in the development of schizophrenia by providing evidence of a causal relation between risk factors and the development of schizophrenia-related behavioural abnormalities and endophenotypes.

Many animal models have been focused on rodents, using mice and rats, whereas others have used nonhuman primates. Given the homology between the human and nonhuman primate brain, the latter models promise to be an important area of research.2,3 However, examination of the genetic mechanisms implicated in schizophrenia is difficult in nonhuman primate models, whereas genetic knockouts can be designed in mouse models with relative ease. In addition, neurodevelopmental processes can also easily be examined in rats. Hence, this review focuses on nonprimate (i.e., rodent) models of schizophrenia, and for the purposes of this review we refer to such models as animal models. Whereas the homology between human and rodent brains is less than that between human and nonhuman primate brains, important endophenotypes and behaviours related to schizophrenia can be examined effectively in rodent models. Descriptions of the homology between human and rodent brains, however, are beyond the scope of this review.

After discussions of the behavioural and biologic features of schizophrenia and the etiology of the disorder, this review discusses the concept of endophenotypes, which serve as a behavioural link between animal behaviour and human symptomatology. This serves as background for a discussion on animal models in schizophrenia research that focuses on animal models of disrupted neurotransmission, environmental risk factors, disrupted neurodevelopment and genetic risk factors. We focus on the behavioural endophenotypes of animal models that are created by particular experimental manipulations (i.e., gene mutations, lesions or drug injections) as a method of understanding etiologic and mechanistic hypotheses of schizophrenia. We argue that animal model research can complement the literature on schizophrenia in humans by using experimental paradigms to test causative and mechanistic hypotheses of schizophrenia in ways that would not be feasible in human participants.

Behavioural and biologic features of schizophrenia

Symptomatology and course

The symptomatology of schizophrenia is generally grouped into positive, negative and cognitive symptoms. Positive symptoms reflect an excess or a distortion in normal functioning and include conceptual disorganization, hallucinations and unusual thought content.1,4 Negative symptoms reflect a loss or diminution of normal functioning and include restricted affect, alogia, anhedonia, decreased sense of purpose, diminished social drive and motor abnormalities.1,4,5 Cognitive symptoms, considered by some to be the most important features of schizophrenia, include impairments in attention and information processing speed, visual and verbal learning, working memory, social learning and executive function.6,7

The active phase of the disorder is generally marked by the emergence of positive symptoms and typically occurs during adolescence, possibly coinciding with the maturation of and axonal pruning in certain brain areas.8 A prodromal phase usually precedes the active phase and presents as a nonspecific behavioural change, although the lack of specificity and predictive validity of the associated symptoms do not permit their use in early diagnosis or treatment.1,8 A premorbid stage may also precede the prodromal stage and may manifest as motor, cognitive and social abnormalities in children in whom schizophrenia later develops.8,9 Should the positive symptoms remit, either spontaneously or as a result of treatment, the individual is said to be in the residual stage, which is primarily characterized by negative and cognitive symptoms.1,7,8

Neurotransmitter system abnormalities

Abnormalities of the major neurotransmitter systems have been reported in the brains of individuals with schizophrenia. Dopaminergic hyperactivity has been reported in the striatum and may be the result of a greater number of D2 receptors that have a higher affinity for dopamine (DA) or an enhanced presynaptic accumulation of DA in the striatum.1012 In addition, dopaminergic hypoactivity has been reported in the prefrontal cortex (PFC).13,14

Glutamatergic hypoactivity has been reported in the brains of patients with schizophrenia. Depletion of glutamate and its synthesizing enzyme, as well as N-methyl-d-aspartate (NMDA) receptors for glutamate, may be responsible for the hypoactivity of the glutamatergic system.1517 Consistent with this, glutamatergic antagonists have been shown to induce acute psychotic reactions, including thought disorder, social withdrawal and catatonia in unaffected volunteers.18

Hypoactivity of the GABAergic system has also been reported in the brains of patients with schizophrenia. This dysregulation is thought to be the result of reductions in key γ-aminobutyric acid (GABA)–related compounds, including glutamic acid decarboxylase (GAD), an enzyme required for the synthesis of GABA, and GABA transporter 1 (GAT-1) synthesis, a presynaptic GABA reuptake transporter.1921

Neuroanatomic brain abnormalities

There are also gross neuroanatomic abnormalities in the brains of patients with schizophrenia, affecting most of the brain’s major structures. Ventricular enlargement, particularly in the lateral and third ventricles, as well as a reduction of cerebral volume have been reported in the brains of patients with schizophrenia compared with controls.2224

Abnormalities of the PFC have also been reported in schizophrenia. These abnormalities include reductions in cortical grey matter, a decrease in the size of pyramidal neurons and a reduction in the total volume of dendrites and axons.25,26 Developmental errors in neuronal migratory patterns have also been observed in postmortem tissue taken from the PFCs of patients with schizophrenia.27,28

Abnormalities of the temporal lobe include reduced grey and white matter volume in patients with schizophrenia.29,30 Volumetric decreases in the superior temporal gyrus, the parahippocampal gyrus, the hippocampus and the amygdala have also been identified.3133 Volumetric decreases of both the hippocampus and amygdala have also been found in patients with schizophrenia, although with some inconsistency.3436 Decreased neuronal size and density, as well as a disorganization of hippocampal neurons, have been reported in the hippocampus.37,38

Volumetric reductions of the thalamus and structures of the basal ganglia in the brains of patients with schizophrenia have also been reported.22,39,40

Etiology of schizophrenia

Evidence of behavioural and cognitive disturbances before the onset of schizophrenia is important because it suggests that the causes of the disorder precede the development of schizophrenia by many years. These causes are thought to include genetic factors, as well as a number of prenatal and perinatal developmental insults. The neurodevelopmental hypothesis of schizophrenia posits that schizophrenia may result from subtle abnormalities affecting critical circuits in the brain during early development, with full-blown consequences becoming evident during early adulthood when the damaged structures become fully functional.9 According to Weinberger,41 if some form of damage affects a structure or region that has not fully matured, the effects of the damage would remain clinically silent until the structure fully develops. However, it is possible that some manifestations of the damage that do not reach clinical significance might be evident in childhood.9 Although Weinberger’s hypothesis suggests that the brain damage results from prenatal and perinatal developmental events, it does not preclude the possibility that the damage might be caused by disrupted genes that begin to exert their effect early in life. In fact, studies have shown a substantial genetic component to schizophrenia, with a heritability (i.e., the proportion of variance in the phenotype that can be attributed to genetic variance) of the disorder of about 80%.6,42

Family, twin and adoption studies

Family studies, in which prevalence data are gathered on parents, offspring, siblings and extended family members, have shown that schizophrenia occurs in about 10%–12% of first-degree relatives and 3%–4% of second-degree relatives of patients with the disorder (i.e., probands), compared with a 1% incidence rate in the general population.4345 However, genetic and environmental effects cannot be well distinguished, as family members often share a similar environment.

Twin studies can be used to address this confound as monozygotic (MZ) twins share 100% of their genes, whereas dizygotic (DZ) twins share on average 50% of their genes. A greater similarity or concordance between MZ twins than between DZ twins suggests a genetic variation underlying the disorder, which has been confirmed by several studies.42,46,47 However, interpretation of these results requires an assumption of the equality of the prenatal, perinatal and postnatal environments, which may not always be the case.46

To address the limitations of twin studies, adoption studies have been used. In such studies, the development of schizophrenia in the adoptee is correlated with the development of the disorder in biologic and adoptive parents and relatives. Adoption studies have consistently shown that the biologic relatives of adoptees with schizophrenia have higher rates of schizophrenia and schizophrenia-spectrum disorders than the adoptive relatives.43,48,49 Although these results provide the best evidence for a genetic component to the disorder, adoption studies do not separate in utero or perinatal environmental effects from genetic effects.

Despite some limitations, family, twin and adoption studies have successfully demonstrated that schizophrenia is not a single-gene disorder, nor is it a collection of single-gene disorders.50 In such models, the penetrance of the disorder (i.e., the probability of the expression of the phenotype given the presence of a particular gene), based on concordance rates between MZ twins, would be 50%, and a linear decrease in risk would be predicted as relatives become more distant, which, as previously mentioned, is not the case.44 Hence, it has been suggested that the mode of inheritance must follow a polygenic (i.e., a large number of genes, each of a small effect) model of inheritance.50

Multifactorial threshold model of inheritance

Most current etiologic theories posit a multifactorial threshold model of inheritance.51,52 In this model, a large number of polygenes and nonshared environmental experiences have interchangeable and additive effects on the risk for schizophrenia.44,51 In addition, the model posits an arbitrary categorical threshold on the dimension of risk, beyond which schizophrenia would develop in an individual.44,51

Multifactorial models posit a greater number of possible risk factors than the number necessary to cross the threshold.44 Hence, individuals in whom schizophrenia develops need only be exposed to a subset of possible risk factors to cross the threshold, which allows for genetic and phenotypic variation among patients with schizophrenia and is consistent with the presentation of the disorder. For example, it has been suggested that paranoid schizophrenia, a less severe form of the disorder, may develop in individuals with fewer polygenes, whereas nonparanoid schizophrenia, a more severe form, may develop in individuals with a greater number of polygenes.53 Multifactorial models also predict a curvilinear decrease in risk as relatives become more distant, which is consistent with family studies.44

Endophenotypes

Clearly, the etiology and symptom presentation of schizophrenia is complex. As this complexity is difficult to replicate in animal models of schizophrenia, particularly in terms of the direct assessment of symptoms, other behavioural markers or biomarkers are necessary to assess for schizophrenia-related symptoms in animals. The concept of a biomarker is used across multiple scientific disciplines. In this context, it is considered to be endogenous and measurable characteristics that indicate the risk for or presence of a psychiatric illness.54 One subtype of biomarkers is the endophenotype, which is more restrictive in its definition.54 Endophenotypes serve as the “bridge” between animal models of schizophrenia and the human disorder. Since actual schizophrenia symptoms generally cannot be directly assessed in animal models, endophenotypes can be used to measure more “upstream” behavioural disturbances that are related to schizophrenia symptoms. Endophenotypes are defined as measurable phenotypes that are unseen by the unaided eye and lie along the pathway between the genotype and the disease.55 In essence, each genetic abnormality would be reflected in a specific protein change, which would be reflected in a discrete functional abnormality, such that an endophenotype can be viewed as the direct phenotype of the abnormal gene.56 It should be noted that abnormal genes can be influenced by multiple factors, including environmental, epigenetic and genetic interactions.55,56 It is also possible that each gene or genetic interaction could give rise to 1 or more endophenotypes, and that the endophenotypes resulting from an abnormal gene may be similar or distinct, depending on where these genes are expressed in the brain.56,57 Endophenotypes can be diverse and may include behavioural, biochemical, neuroanatomic, cognitive or endocrinologic measures.55 To gain further understanding of the etiology and underlying mechanisms of the symptomatology of schizophrenia, we mainly discuss behavioural endophenotypes as they relate to particular animal models. We also discuss neuroanatomic and neurotransmitter abnormalities as appropriate.

There are several guidelines used to identify endophenotypes, which render them more specific than general biomarkers:55,56

  • the endophenotype is associated with the illness;

  • the endophenotype is heritable, thereby implying a genetic basis;

  • the endophenotype is state-independent, such that state-related changes in the individual’s status (i.e., remission) should not affect the expression of the endophenotype;

  • endophenotypes cosegregate with the illness within families; and

  • the proband’s endophenotype is found at a higher frequency in the proband’s relatives than in the general population.

Consistent with the last point, since schizophrenia consists of multifactorial traits and each individual trait segregates independently in family members, some unaffected relatives will express some endophenotypes linked with the disorder, but others will not.56

According to this approach, schizophrenia symptoms are thought to result from the combination of multiple endophenotypes. Thus, defective genes do not code for the disease directly, but rather for physiologic processes that culminate “downstream” in the development of the disease.57,58 Different combinations of risk factors could therefore give rise to different combinations of endophenotypes. Consistent with the heterogeneity of schizophrenia, these differential endophenotype combinations may also result in differential symptom presentation. For example, different genetic risk factors may distinguish between paranoid and nonparanoid schizophrenia subtypes.53

One major advantage of using endophenotypes as behavioural markers is that they are particularly useful when the imprecision of psychiatric diagnoses can impede genetic investigations.59 Specifically, the relation between the endophenotypes and particular genes should show a stronger association than that between risk genes and the symptoms of schizophrenia.56 It should be noted that the standard definition of an endophenotype requires that it be heritable or genetically based. However, in many animal models, environmental manipulations are used to mimic some form of brain damage, and the behavioural features of endophenotypes associated with schizophrenia are examined. In addition, in most genetically based animal model studies, many of the criteria required to meet the definition of an endophenotype are not met or remain untested. For example, in some studies the heritability or the state independence of the genetic manipulation are not measured. Although the results of certain assays are often referred to as “endophenotypes” in the literature, they are more correctly defined as biomarkers if all the criteria of an endophenotype are not met or tested. Hence, when presenting results of animal model studies, we either refer to the observed behavioural abnormalities as such, or as biomarkers.

Multiple endophenotypes have been associated with schizophrenia, and similar biomarkers have been assessed in animal models of the disorder. Only those that are commonly measured in rodents will be discussed in detail in subsequent sections. However, it is worth briefly mentioning several endophenotypes that are based on research with human patients because they could theoretically be assessed in animal models, more likely in primate than in rodent models. It should be noted that many of the endophenotypes and biomarkers that we discuss are not exclusive to schizophrenia and are associated with other psychiatric disorders.

Deficits in smooth pursuit eye movements have been among the most reliable biologic findings in schizophrenia research and may be representative of frontal lobe dysfunction and deficient stimulus encoding processes.60,61 Several studies have reported abnormalities of smooth pursuit eye movements in both patients and their relatives.62,63 A second human-related endophenotype is impairment in the P300 response, which is an event-related potential that is thought to represent brain activity resulting from tasks that require information to be maintained in working memory.64 Several meta-analyses have implicated the P300 response as an endophenotype of schizophrenia, and studies have shown that unaffected family members of probands exhibit deficits in the P300 response.65,66 Deficits in P50 suppression, which is operationalized as a decrease in the amplitude of the P50 wave (i.e., a positive-going wave at 50 ms latency) to the second of 2 paired auditory stimuli, is thought to reflect a sensory gating mechanism.65 Abnormalities in P50 suppression have been described in both patients with schizophrenia and their unaffected relatives.65,67 Finally, episodic memory retrieval, which is defined as the conscious recollection of an event by reliving it mentally, appears to be disrupted in patients with schizophrenia, irrespective of medication status, and in unaffected relatives of schizophrenia probands.68,69

Sensorimotor gating deficits

Sensorimotor gating is a process whereby excess or trivial information is screened out of awareness (i.e., gated out), permitting the individual to focus on the more important stimuli in the environment.70 Deficits in the ability to filter irrelevant internal and external stimuli may result in misperceptions, a sense of sensory flooding and disorganized thinking and distractibility, which are all reminiscent of the positive symptoms of schizophrenia.71 In fact, studies of precategorical processing have suggested that patients with schizophrenia have deficits in selective attention or a “filtering deficit.”72

Prepulse inhibition (PPI) of startle is a commonly used measure to test sensorimotor gating in humans and animals.73 The procedure is based on the fact that a brief, startling stimulus will produce a startle response. If a weaker, nonstartling prepulse precedes the startling stimulus by a short time interval (i.e., 30–300 ms), the startle response will be reduced.74 This reduction is thought to result from a momentary inhibitory sensorimotor gating process that is caused by the non-startling prepulse, which serves to protect its ability to be processed.74

Several studies have reported that patients with first-episode schizophrenia and medicated patients with acute psychosis have deficits in acoustic PPI, especially when strong prepulses are used.71,75,76 Using an eye-blink PPI paradigm, it has been shown that patients with schizophrenia, their unaffected relatives, and individuals with schizotypal personality disorder had deficits in right-eyeblink PPI compared with controls.77 Studies have also indicated that deficits in PPI often correlate with both positive and negative symptoms.74,75,78 However, despite being one of the most replicable findings in schizophrenia, it is important to note that a deficit in PPI is not exclusive to schizophrenia and has been observed in many other psychiatric disorders.79

Deficits in sensorimotor gating can be observed in animal models of schizophrenia using the PPI test with mice or rats. The animal is first placed in a small restraining device to minimize movement and is then placed in an apparatus that can deliver acoustic startle pulses and prepulses. A sensor records startle responses to startle-alone trials and to prepulse + startle trials. The relative difference between the 2 startle responses constitutes PPI. Rats treated with pharmacologic agents and environmental insults, as well as genetic mutant mice, have all shown deficits in PPI.79,80

Working memory deficits

Working memory is conceptualized as a “limited capacity storage system, which temporarily maintains and stores information [and] supports human thought processes by providing an interface between perception, long-term memory and action.”81 Many studies have found impairments in working memory in patients with schizophrenia and their unaffected siblings. Studies have suggested that patients with schizophrenia may have deficiencies in active rehearsal processes, whereas passive stimulus maintenance processes and long-term memory networks may remain unaffected.72

More specifically, spatial working memory may be one possible endophenotype of schizophrenia. Some studies have shown that patients with schizophrenia and their unaffected relatives performed worse than controls on spatial working memory tasks, but that only patients performed significantly below controls on working memory tasks across other domains (i.e., verbal and object).8284 However, another study has suggested that spatial working memory deficits were only present in family members with diagnosable schizophrenia-spectrum personality disorders.85 Other studies have also suggested that patients with schizophrenia may have a reduced memory span across modalities.86,87 Two meta-analyses that examined working memory in patients with schizophrenia have suggested that the latter showed a substantial effect size for working memory deficits, regardless of modality, although there may be more consistent impairments in visual working memory.88,89 However, it also has been suggested that working memory deficits may be the result of a general encoding deficit, particularly stimulus encoding (i.e., the transformation and preparation of presenting stimuli into a format facilitating collateral functions, including those of “working memory”).9092 In addition, it has been suggested that impairments in working memory may be the result of deficits in processing speed.93 Moreover, developments in mathematical modelling of schizophrenia cognition have formally integrated memory-trace impairment and encoding elongation.92

Spatial working memory is easily assessed in animal models of schizophrenia using spatial memory tasks. The Morris water maze is the most commonly used task in behavioural neuroscience research and can easily be configured to test spatial working memory with rats.94 In this procedure, rats are placed in a pool of opaque water where there is a hidden platform located slightly below the water’s surface. The rat is then released into the water at various points around the periphery of the pool and must navigate to the platform on the basis of the spatial cues in the room.94 Although the water maze protocol can be used to test working memory in mice, this species tends to float and does not remain on the platform once it is reached, making interpretation of its behaviour uncertain.95 Hence, the paddle pool task was developed to remove these confounds.95 In this task, the mouse is placed in a circular pool that contains water to a depth of 2 cm. This motivates the mouse to escape but does not require it to swim. There are 12 exit holes on the wall of the pool, only 1 of which leads to an actual exit, which is connected to the mouse’s home cage. The mouse is released near the centre of the pool and must learn to find the exit hole based on the spatial cues in the room.95

Although these tasks require the acquisition of a spatial working memory, simple versions of the tasks do not unambiguously measure working memory. Therefore, special versions of the tasks have been used to assess spatial working memory. These typically include a period of initial acclimation and training in a conventional version of the task, followed by frequent reversals of the hidden platform or exit location or new matching-to-place tasks on successive days.96,97 These versions engage frontal cortex mechanisms by using retention intervals in the working memory time-frame and also allow for intertrial proactive interference.98 Other tests of working memory include tests of continuous delayed alternation, discrete paired trial variable-delay alternation tasks and the radial arm maze.99 Many animal models, including pharmacologic, neurodevelopmental and genetic models, have shown impairments in these working memory tasks.99

Stereotypy and perseverative behaviour

Stereotypy and perseverative behaviour are also considered to be viable endophenotypes, as well as symptoms, of schizophrenia, and are related to working memory performance. For example, the ability to set-shift is necessary for successful performance of the modified versions of the Morris water maze. In fact, studies have shown that patients with schizophrenia are often impaired on both reversal learning and extra-dimensional set-shifting.100,101 Studies have shown that patients engage in perseverative behaviour on the Wisconsin Card Sorting Test (WCST), which is considered to be a measure of prefrontal cortical activity, as well as on a measure designed specifically to assess stereotypy, the Stereotypy Test Apparatus.102,103 Studies have also shown that first-degree relatives of schizophrenia probands demonstrate perseverative behaviour on the WCST, further suggesting its relevance as a measureable endophenotype.104,105

Assessing perseverative behaviour is relatively simple in animal models. Rats given psychostimulants often exhibit locomotor hyperactivity, and at higher doses they exhibit stereotyped or perseverative behaviour.106 Although people with schizophrenia generally do not exhibit hyperlocomotion, they do often engage in stereotyped or perseverative behaviour, which is thought to be modelled by stereotyped behaviour in rats.106 Hyperlocomotion and stereotyped behaviour are often seen in pharmacologic animal models and are thought to model the positive symptoms of schizophrenia.107,108

Deficits in latent inhibition

Latent inhibition (LI) is a process whereby an unreinforced stimulus interferes with the conditioning of a new contingency to the same previously unreinforced stimulus.109 In such studies, the conditioned stimulus (CS) is presented alone (i.e., the CS-nothing contingency). Later, the CS is paired with an unconditioned stimulus (US), usually an aversive stimulus (i.e., the CS–US contingency). The CS-nothing pairing is thought to proactively interfere with the learning of the CS–US pairing, such that the conditioned response (CR) to the CS is reduced.109 This reduction in the CR to the CS is known as LI. The interference in learning the CS–US contingency is thought to result from a decline in attentional processing of the CS when it was presented alone, such that the absence of LI is interpreted as an inability to discern relevant from irrelevant stimuli.110

Several studies have shown that LI is disrupted in patients with acute schizophrenia, but not those with chronic schizophrenia, although 1 study has found LI to be disrupted and persistent in the latter population.110112 These differences noted between patients with acute and chronic schizophrenia do not appear to be the result of stabilization due to medication, but rather the result of an evolution of the intrinsic factors of schizophrenia.112 Impairments in LI have been found in first-degree relatives of schizophrenia probands, regardless of whether these relatives displayed schizotypal features.113,114 One study, however, has not found an association between LI and schizophrenia.115 Latent inhibition is thought to correlate with both positive and negative symptoms, depending on the methodology used.109

Latent inhibition tasks are easily conducted with rats and mice. Generally the CS is a tone, the US is a footshock, and the CR is a freezing response (i.e., conditioned fear). The differential response between the presentation of the CS–US contingency without the prior presentation of the CS alone and the presentation of the CS–US contingency with the prior presentation of the CS alone is an index of LI. Abnormal LI has been observed in pharmacologic, neurodevelopmental and genetic animal models of schizophrenia.109

Social withdrawal

Although not strictly considered to be an endophenotype, but rather a negative symptom of schizophrenia, social withdrawal is often assessed in animal models of schizophrenia. Studies have shown that adult patients with schizophrenia have a reduced social network and deficits in social competence, social skills and social cognition, including the misperception of affective information.116119 Social anhedonia has also been observed in the relatives of schizophrenia probands.120,121

Rodents are typically social animals and tend to approach unfamiliar conspecifics. As such, the social interaction test in animal models of schizophrenia is designed to measure social impairments similar to those in human patients. Recent developments in automated, digitized equipment and software have allowed the efficient collection of data on the social behaviour of pairs of animals on a time scale of seconds to weeks.107,122 Tests can include observations of animal dyads in their home cages122 or in a novel arena107 or while 1 animal’s movements are limited by placing it under a wire cup.123 Regardless of the test used when studying social behaviour, the basic expectation is a reduction in social behaviour. Pharmacologic, genetic and neurodevelopmental animal models of schizophrenia have consistently found a reduction in social behaviour.107,122,124,125

Animal models in schizophrenia research

Animal models of schizophrenia can be developed in several different ways. Pharmacologic animal models are often used to test hypotheses related to glutamatergic and dopaminergic dysfunction. Neurodevelopmental models are used to either mimic certain neurodevelopmental risk factors of schizophrenia or to heuristically approximate damage that might occur by other means in schizophrenia. Gene knockout mouse models of schizophrenia target and disrupt specific genes that are thought to be risk factors in schizophrenia.

Ideally, an animal model should begin with the known pathogenesis of the disease, such that modelling schizophrenia risk factors in rodents would increase the value of these models.126,127 Given that the etiology and symptomatology of schizophrenia is complex, no animal model will be able to capture its full complexity, particularly since the full complexity of human etiologic interactions and most symptoms cannot be modelled in rodents. Delusions, hallucinations, disorganized thinking, affective flattening, alogia and avolition are virtually impossible to model in animals, although it has recently been suggested that impaired reality testing may be modelled in rodents using certain Pavlovian conditioning procedures.128,129 Given that schizophrenia is often characterized as a higher-order cognitive disorder, the ability to faithfully model symptoms in less cognitively developed animals can be difficult.130 As a result, researchers have had to rely on the assessment of biomarkers and endophenotypes. However, despite being present in patients with schizophrenia, the relation between these biomarkers and the symptoms is not always clear, and many biomarkers are also observable in other disorders.128 For example, social interaction deficits are observable in animal models of schizophrenia and autism.107,131 However, the necessity of using endophenotypes and biomarkers in animal models of schizophrenia might also be an advantage of such studies, given the closer relation between risk factors and endophenotypes compared with the more distant relations between risk factors and symptoms. As previously mentioned, human association studies may also benefit from the use of endophenotypes as phenotypic markers. In addition, the lack of complete homology between the rodent and human central nervous systems (CNS) and the unknown specifics of how a particular endophenotype is expressed in each species mean that animal model data must be related back to schizophrenia with caution. It should also be noted that there is a great deal of behavioural variability in different mouse strains used to generate animal models, thereby requiring additional caution in interpreting behavioural biomarkers.132

Despite these shortcomings, animal models are not meant to serve as complete equivalents to the disorder. Rather, animal models are useful in testing causative or mechanistic hypotheses of schizophrenia.130 Animal models, while not providing a complete account of schizophrenia, complement the literature on schizophrenia in humans, which is largely correlational, by their ability to experimentally manipulate and control factors that could not otherwise be manipulated or controlled in humans. This permits a causal determination of the relation between the manipulation, whether it be pharmacologic, neurodevelopmental or genetic, and the presence of behavioural biomarkers.

Animal models can also be useful in understanding the mechanisms underlying the development of schizophrenia. In both genetic and neurodevelopmental models, brain tissue can be examined at different points throughout the lifespan to determine how the disease progresses and what factors may be involved in its progression. For example, the dysregulation of subcortical DA may permit a determination of specific brain changes that result from chronic dopaminergic dysregulation and of how those changes might impact other areas of the brain.

The fact that animal models often involve 1 particular gene knockout or environmental insult allow behavioural abnormalities associated with particular risk factors to be assessed without confounds from other risk factors that would be expected given the multifactorial nature of schizophrenia. This may also provide a better understanding of the etiology and underlying mechanisms of schizophrenia as they relate to a particular risk factor. In a sense, each model may provide 1 piece of the puzzle in understanding the causal relations between the etiology and mechanisms of schizophrenia and the development of the disorder. Although gene–environment interaction models can be created and would be extremely useful in advancing the field, such models are rare.133 Hence, a single animal model cannot represent the entire population of patients with schizophrenia, but rather only a subset of patients with a particular risk factor.126 When taken together, the resulting information can provide a more complete understanding of schizophrenia in a way that would not be feasible in studies on schizophrenia in humans.

Animal models of neurotransmitter dysregulation

As previously mentioned, studies in schizophrenia in humans show a dysregulation in several neurotransmitter systems, including those involving DA, glutamate and GABA. Animal model studies have typically focused on dopaminergic, glutamatergic and, to a lesser extent, GABAergic transmission, either through pharmacologic or genetic manipulations.

Pharmacologic animal models

Pharmacologic animal models typically involve the injection of amphetamine, a dopaminergic agonist, or phencyclidine (PCP), ketamine or MK-801, which are glutamatergic antagonists, into a rat or mouse.

As previously mentioned, patients with schizophrenia show dopaminergic hyperactivity in the striatum and dopaminergic hypoactivity in the PFC.10,13 In animal models, administration of chronic amphetamine by implantation of slow-release silicone pellets is thought to model paranoid schizophrenia.134 As time progresses, rats show increased locomotion (0–6 h), motor stereotypies (6 h to 3 d), reclusion to their burrows (4 d) and aggressive social interaction (5–7 d).134 Amphetamine administration has been shown to disrupt PPI, spatial learning and social interaction.135137 Interestingly, other studies have found no effect of amphetamine on social interaction, but noted an increase in stereotypy and locomotor activity.107,108 Dopaminergic antagonists applied to the PFC have been shown to disrupt PPI but not LI.138

As previously described, patients with schizophrenia show glutamatergic hypoactivity, such that the administration of glutamatergic antagonists induce acute psychotic reactions, including thought disorder, social withdrawal and catatonia in normal volunteers.18 In animal models, gultamatergic antagonists, such as PCP, disrupt PPI, spatial learning and social interaction.135,137,139141 Increased locomotion and stereotyped behaviour were also evident in rats treated with single and repeated doses of PCP.107,142,143 The social interaction deficits and stereotypy caused by PCP administration were reversed by conventional and novel antipsychotic medications.144,145 The administration of ketamine or MK-801 also disrupts LI, spatial learning and social behaviour.135,146,147 In fact, early repeated blockade of NMDA receptors using MK-801, where injections were administered from postnatal days 6–21, showed behavioural deficits, including decreased locomotion and exploratory behaviour, in adulthood.148 This latter study better approximates glutamatergic dysfunction in schizophrenia, as the disruptions were made both repeatedly and early, whereas other studies merely used 1 acute injection. Consistent with the neurodevelopmental hypothesis, any disruption should occur early in life. Hence, the acute versus perinatal administration of glutamatergic antagonists may be responsible for the behavioural differences observed among studies.

As previously mentioned, the brains of patients with schizophrenia also show hypoactivity of the GABAergic system. The administration of picrotoxin, an antagonist of GABAA receptors, in the medial PFC but not in the lateral PFC or the hippocampus, has been shown to reduce PPI in rats.149 Furthermore, it has been shown that blockade of GABAA receptors in the rat amygdala leads to abnormal GABAergic transmission in the hippocampus in a manner that is consistent with patients with schizophrenia.150152

Taken together, the results show that pharmacologic animal models are useful in understanding the behavioural features associated with widespread neurotransmitter dysfunction, although the specific mechanisms are not fully known.153,154 However, these models have limited use in assessing the developmental nature of the disorder. It has been suggested that the predictive and explanatory capabilities of pharmacologic models have, in fact, reached their limit.130

One point of interest is that many of these pharmacologic manipulations can be conducted in humans with and without schizophrenia. Hence, their use in complementing the literature on schizophrenia in humans is limited. However, they are useful in demonstrating that similar biomarkers are observable in both humans and rodents when specific pharmacologic agents are administered. The identification and the expression of these biomarkers can then be assessed in other animal models.

Another shortcoming of pharmacologic models is that the transient nature of the drug effects does not follow an etiologically relevant course, as no permanent changes in the brain are expected. In addition, given the neurodevelopmental nature of schizophrenia, early-life neurotransmitter dysfunction would be expected to have an impact on brain development. Such an impact would not be expected in this type of animal model, as the animals will have undergone normal development before drug administration. Even when chronic drug administration occurs, prenatal and early perinatal neurotransmission function is normal. However, downstream effects of this neurotransmitter dysregulation can be examined in these models, without any confounds from morphologic brain abnormalities or other dysregulated neurotransmitter systems, both of which could be expected in patients with schizophrenia.

Genes involved in neurotransmission

Genetic animal models based on neurotransmitter abnormalities have focused on the disruption of the dopaminergic and glutamatergic systems. Dopamine transporter (Dat) knockout mice are unable to reuptake released DA, resulting in a hyperactivity of the dopaminergic system.155 Similar to the findings in rats treated with amphetamine, Dat knockout mice display hyperactivity, perseverative locomotion and deficits in PPI and spatial learning but no deficits in social behaviour.155157

The NMDA receptor is one of the receptors for glutamate, and mutations have been made in the NR1 subunit of the receptor such that the receptor’s functionality was lowered.155 Similar to the findings of PCP administration in rats, mice with NMDA receptor NR1 subunit hypofunction display deficits in PPI, hyperactivity, decreased anxiety-related behaviours and deficits in social interaction.158160 Calcineurin may be involved in the glutamatergic system and in the neurodevelopment of the brain.161 Conditional calcineurin knockout mice displayed hyperactivity, deficits in PPI and LI, as well as abnormal social behaviour over a 3-day period.122

Taken together, these genetic animal models tend to show similar results to their pharmacologic counterparts. However, given the persistent developmental disruption of these genes, genetic models provide a greater consistency with the etiology of schizophrenia. As a result, they appear to be more informative than the aforementioned pharmacologic models in terms of etiology and underlying developmental mechanisms.

Animal models of environmental risk factors

Environmental risk factors in the multifactorial threshold model are generally considered to be those occurring prenatally or perinatally, although postnatal environmental stressors may also contribute to risk.

Viral exposure

In utero viral infections are examples of prenatal factors that may increase the risk of schizophrenia. In utero exposure to maternal influenza has been implicated in increased risk for schizophrenia.162,163 In addition, prenatal exposure to rubella, toxoplasmosis and herpes simplex virus type 2 have all been linked with the development of schizophrenia.164

In animal model studies, maternal exposure to influenza was found to upregulate several genes and downregulate others that are involved in signal transduction, transport, protein metabolism and cell growth, as well as some genes that have been implicated in schizophrenia.165,166 Studies have also shown that maternal influenza infection results in deficits in PPI, social interaction and exploratory behaviour.167169 Exposure to the herpes simplex virus and the Borna disease virus also impaired PPI.170,171 However, it has been suggested that it is not fetal exposure to the virus itself that confers risk but rather exposure to the maternal immune response that may lead to altered brain development.166,167,172,173 In fact, maternal exposure to a viral mimic, polyinosinic:polycytidylic acid (Poly I:C) or lipopolysaccharide (LPS) leads to deficits in LI, working memory, avoidance learning and PPI.167,174,175 Although the precise effects of the maternal immune response on the developing brain are unknown, several hypotheses have been advanced:

  • the maternal immune response elevates cytokines in the placenta, amniotic fluid and in the fetal brain, resulting in an inflammatory reaction that may impact neurodevelopment, fetal growth and placental function;

  • fever, as a consequence of cytokine release, can result in abnormalities of the CNS and apoptosis in the cerebral cortex; and

  • antibodies resulting from the infection may react and injure fetal brain structures.168,176

Obstetric complications

Perinatal obstetric complications have been found to increase the risk of schizophrenia developing later in life.9 It has been shown that children born with obstetric complications at delivery, including preeclampsia, gestational age of less than 33 weeks, inertia of labour, vacuum extraction and respiratory illness, have about twice the risk for schizophrenia than those without such complications.177,178 Whereas these obstetric complications are not specific to increasing risk for schizophrenia, they have been hypothesized to be etiologically relevant in the presence of schizophrenia risk genes.179 Since the expression of many genes typically changes throughout development, a genetic predisposition involving the lack or early expression of susceptibility genes may cause an individual to be more susceptible to perinatal insults.9,180 In fact, it has been shown that several schizophrenia risk genes that are regulated by hypoxia or involved in vascular function in the brain showed a significant gene × obstetric complications interaction.181 This suggests that mutations in susceptibility genes may render schizophrenia more likely to develop in the presence of an obstetric complication. Similarly, it has also been suggested that in the presence of genetic vulnerability, postnatal stressors can result in the onset of or an increase in schizophrenia symptomatology.182

Several animal models of obstetric complications have been assessed for schizophrenia-related biomarkers. Fetal hypoxia has been shown to lead to schizophrenia-related abnormalities in rats.183 Rats exposed to neonatal asphyxia by being placed in a chamber of nitrogen gas for 30 minutes, showed amphetamine-induced hyperactivity, stereotypy, decreased social interaction and decreased brain-derived neurotrophic factor, the latter a finding in some patients with schizophrenia.184186 Another study has shown that 3 months after neonatal asphyxia, rats had an increase of vesicular monoamine transporter, which is involved in DA transport, in the striatum and an increase of glutamate transporter in the frontal cortex.187

Prenatal stress has also been used in the development of animal models of schizophrenia. In 1 study using maternal restraint stress, DA and glutamate receptors in the offspring were increased in the dorsal frontal cortex, the medial frontal cortex and the hippocampus.188,189 Early maternal deprivation effects on offspring have included retarded motor development, reduced locomotion in an open field, increased amphetamine-induced locomotion and impairments in PPI and spatial memory.190,191

The aforementioned animal models replicate actual neurodevelopmental insults that are correlationally related to the development of schizophrenia and are useful in understanding the direct impact of prenatal and perinatal insults in the development of schizophrenia. Since these models replicate actual risk factors associated with schizophrenia, they would be most useful in assessing the neurodevelopmental nature of the disorder. Whereas the literature on schizophrenia in humans has provided correlational evidence between these insults and the development of schizophrenia, these animal models can address the question of a causal link between the risk factor and the development of specific schizophrenia-related biomarkers. These models are also useful in characterizing the mechanisms by which these insults exert their effects. For example, given the nature of human prenatal viral exposure research, it would be difficult to determine that it is the immune response, rather than the virus itself, that confers risk for the development of schizophrenia. Clearly, given the extreme manipulations described above, no such experiments could ever be conducted on human participants. Hence, the specific relations between the neurodevelopmental insults and schizophrenia-associated behavioural abnormalities can be elucidated mainly through animal model research. This kind of research exemplifies the relation between the human and animal literature. It is unfortunate that a full behavioural characterization of many of these models has not yet been conducted. However, given their strong etiologic relevance, future research into these models holds promise for further understanding the etiology of schizophrenia.

Animal models of neurodevelopmental disruption

Given the neurodevelopmental nature of the disorder, several animal models have attempted to heurisically replicate in animals brain damage that may occur by other means in human patients. Such models permit an understanding of the developmental sequellae that result from particular forms of neurodevelopmental disruptions.

Neonatal ventral hippocampal lesions

Neonatal excitotoxic damage to the ventral hippocampus (VH) is thought to disrupt the development of both the subcortical dopaminergic system and widespread cortical and subcortical circuitry in which the hippocampus participates, including projections to the PFC.130,192 Similarly, inactivation of the VH is thought to disrupt normal maturation of the PFC.192

The appearance of abnormalities in this model follows the neurodevelopmental hypothesis in that they emerge postpubertally.125,193 Studies involving neonatal ibotinic acid lesions to the VH have reported the postpubertal emergence of hyperlocomotion and deficits in PPI, LI, working memory and social interaction.125,193197 Interestingly, these deficits were not evident in rats that received postpubertal lesions of the VH.194,197

Rats neonatally lesioned in the VH showed a postpubertal downregulation of D1 and D2 receptor binding in the striatum, as well as an enhancement of glutamate binding in the PFC, which suggested hyperactivity of the dopaminergic system and hypoactivity of the glutamatergic system.198 In addition, dopaminergic modulation of interneurons in the PFC has been shown to be disrupted.199 It has also been shown that neonatal ventral hippocampal lesions result in a decrease in GAD–producing neurons and a decrease of GABA–related inhibitory interneurons in the hippocampus, enthorinal cortex and PFC.200 Given the presence of multiple behavioural, neurochemical and neuroanatomic abnormalities that are consistent with schizophrenia, this animal model is useful in understanding the effects of suspected dysregulation of neurotransmitter systems in a developmental manner. However, although the resulting abnormalities are thought to arise from prefrontal cortical and dopaminergic disruptions, it is also possible that they may instead arise from indirect results of the lesion or the lesion itself.

The purpose of the VH lesion model is thought to serve as a heuristic to replicate both morphologic abnormalities in the PFC, as well as the dysregulation of neurotransmitter systems in a developmental manner. For example, it has been suggested that, as a direct result of the lesion, maturation of interneurons in the PFC is abnormal, which is consistent with human studies of schizophrenia.199,201 This model is beneficial, as these abnormalities will be present during perinatal stages of brain development in the model, increasing the etiologic relevance. However, it should be noted that, although there are morphologic abnormalities observed in the hippocampus of patients with schizophrenia, the particular damage involved in the animal model (i.e., the complete destruction of the VH) is not observed in human patients. Hence, although the lesioning of the VH may replicate some of the downstream effects of schizophrenia, in terms of behavioural and brain abnormalities, such effects likely arise for different reasons in humans. However, resulting behavioural, neurochemical and neuroanatomic changes are reminiscent of schizophrenia and appear to model appropriate schizophrenia-related biomarkers.

Neonatal medial prefrontal cortex lesion

Given the importance of the PFC in the etiology of schizophrenia, some studies have attempted to lesion the medial PFC (mPFC). The PFC is connected with multiple cortical areas and structures and also modulates some neurotransmitter function, notably the dopaminergic system.202 Studies have shown no effect of the lesion on PPI or locomotor activity; however, there was increased sensitivity to a dopaminergic agonist on PPI in adult animals.202 In addition, it has been found that the lesion, in conjunction with chronic pubertal treatment with a cannabinoid agent, reduces social behaviour and impairs object recognition memory.203,204 Nonbehavioural abnormalities include enlarged ventricles, failure of myelinization of projections from the mPFC to the thalamus, hippocampus, nucleus accumbens and amygdala, as well as an increase in the sensitivity of the mesoaccumbal dopaminergic system.202,205

This model is quite beneficial in that it models developmental deficits in the PFC that are etiologically relevant to the development of schizophrenia. However, as with the VH lesion model, frank lesions of the mPFC are not observed in humans. Given the limited number of studies that have examined this lesion, more work is needed to clarify the nature of resulting biomarkers and the mechanisms underlying the behavioural disturbances. Nevertheless, this model appears to be a promising avenue for further research.

Neurogenesis disruption

One model that involves neurogenesis disruption is the methylazoxymethanol acetate (MAM) model of schizophrenia. In this model, a pregnant dam is treated with MAM, an antimitotic compound that leads to the methylation of nucleic acids and the death of cells that are actively replicating DNA.206 This disrupts brain development in the offspring, particularly neuronal proliferation in the entorhinal cortex.206208 Offspring of MAM–treated dams display hyperlocomotion and rearing; a hyperreactive stress response; dopaminergic hyperresponsivity; and deficits in PPI and LI, passive avoidance learning, object recognition, social interaction and reversal learning.207211 In contrast, some studies have found no deficits in social interaction or spatial and working memory.207,208 Brain changes include abnormalities of the entorhinal, frontal and occipital cortices, the thalamus, the parahippocampal region, the hippocampus and the amygdalohippocampal complex.206,211

Other studies have perturbed epidermal growth factor (EGF) signalling in the brain by administering early postnatal injections of exogenous EGF. In the CNS, EGF enhances the survival and promotes the differentiation of neurons, particularly the dopaminergic neurons of the midbrain.212 Studies have shown that exogenous postnatal injections of EGF lead to hyperlocomotion and increased DA synthesis, and deficits in PPI and social interaction.212,213

Other studies have attempted to perturb neurotrophic signalling. One study infused p75 antibodies conjugated to saporin into the developing PFC, causing loss of the cells in the subplate and the marginal zone.214 The results were the postpubertal emergence of amphetamine-induced locomotion and rearing, and a deficit in PPI. A subsequent study that used postnatal infusions of exogenous nerve growth factor (NGF) into the developing PFC, which caused similar damage as in the p75 model,214,215 also reported similar brain abnormalities consistent with those observed in schizophrenia, as well as hyperactivity at 6 weeks of age and deficits in social interaction in adulthood.216

Taken together, the results of these models are useful in that they permit an understanding of behavioural and brain abnormalities resulting from a disruption of neuronal development and migration in a manner that is consistent with the etiology of schizophrenia. Again, this permits a more specific assessment of mechanisms, behavioural abnormalities and downstream neuroanatomic and neurochemical disruptions, independent of additional genetic or neurodevelopmental risk factors. The results of these models indicate that disruptions in neurodevelopment and cellular migration are causally linked to the development of many schizophrenia-related biomarkers. However, as in the VH lesion model, the causative neurodevelopmental disruption is not etiologically relevant. For example, patients with schizophrenia have not been perinatally injected with EGF or NGF.

A potential problem with these models is that the injection of an exogenous compound may result in nonspecific effects in the brain or other organs. For example, although it has been shown that MAM–exposed offspring display dopaminergic dysregulation and cortical impairments consistent with the neurodevelopmental hypothesis of schizophrenia, the administration of an antimitotic compound could also disrupt functioning in many other organs.217 Despite these shortcomings, such models complement the literature on schizophrenia in humans by providing a more direct assessment of the effects of abnormal neuronal development and cell migration on the development of schizophrenia in a manner that could not be carried out with human participants.

Animal models of genetic risk factors

Given the evidence for a genetic component to schizophrenia, many studies have been undertaken to identify the genes that confer risk for the disorder. As one might expect, based on the brain abnormalities of patients with schizophrenia, many of the implicated genes are those involved in regulating neurodevelopment or neurotransmitter systems. As studies on schizophrenia in humans can only provide correlational analyses, genetic animal models are important in establishing a causative association between a biomarker and a particular risk gene. To describe the complementary nature of both human and animal approaches, We briefly review the methodology and results of both human linkage and association studies and genetic animal model studies. A more detailed discussion of specific genes that have been associated with schizophrenia follows, from both human and animal model perspectives. Since genes specifically related to neurotransmitter dysfunction have already been discussed, we discuss only the other major schizophrenia risk genes here.

Human genetic studies

A linkage analysis is a method for mapping the loci of genes related to a particular disease. The details of the linkage analysis method have been reviewed elsewhere.218 Briefly, the demonstration of nonindependent segregation between schizophrenia and a genetic marker (i.e., another gene or nucleotide sequence) indicates the presence of a disease allele on the chromosome containing that marker.218

Several meta-analyses have been conducted on linkage studies, confirming that schizophrenia is not linked to a single gene locus, but may be linked to multiple genes on multiple chromosomes, as hypothesized by the multifactorial threshold model of inheritance.180,219 Badner and Gershon220 examined all published genome scans for bipolar disorder and schizophrenia in which the location of the marker was provided and significant p values were reported. They found that the strongest linkage with schizophrenia was located on chromosomes 8q, 13q and 22q.220 They also determined that there was some evidence for significant linkage on chromosomes 1q, 2q, 6q and 16q, but that these results were likely owing to the results of single studies.220 Another meta-analysis of 20 genome scans suggested that chromosome 2p12–2q22.1 was significantly linked with schizophrenia when using stringent statistical criteria, whereas when using less stringent statistical criteria, the results showed linkage on chromosomes 5p, 3p, 11q, 6p, 2q, 1p, 22q, 8p, 20p and 14q.221 Another recent linkage analysis study that employed a full genome scan found similar results with linkage on chromosomes 8p, 9q, 8q and 2q.222 One point of interest is that the linkage results are somewhat inconsistent, possibly arising from differences in study selection criteria in the case of the meta-analyses and study methodology, although there are some chromosomal locations that are implicated in multiple studies.

Deletions of one noteworthy chromosomal area, 22q11, are often implicated in the development of schizophrenia. The 22q11 deletion syndrome involves a deletion of the chromosome at 22q11.2 during meiosis.223 Children with this deletion show physical abnormalities, including congenital heart disease, facial deformities and immune system deficiencies, and neuropsychologic deficits, including impairments in visual memory, visual attention, working memory and motor function.224 In addition, children with the 22q11 deletion syndrome are at about 25 times greater risk for schizophrenia developing later in life.225,226 Several studies have also shown that children with the 22q11 deletion syndrome show decreases in PPI.223

Although linkage studies are powerful methods for determining the location of specific risk genes and their locations, they are limited by the fact that they provide only the chromosomal location of risk genes without providing any indication as to which gene is disrupted. To better understand the etiology of schizophrenia, it is crucial to identify specific risk (i.e., disease) genes to elucidate the disruption of biologic pathways.

Human association studies allow for the determination of correlations between particular genes and the development of schizophrenia. The methodology for these studies has been described elsewhere;218 briefly, a particular allele may be considered to confer risk for the development of schizophrenia if the correlation between them is higher than would be expected by chance.218 However, correlation does not necessarily mean causation. In fact, there are several noncausal explanations, such as linkage disequilibrium (i.e., the disease allele is near a gene that has been associated with the disease) or population stratification (i.e., the existence of several subgroups in the population with higher frequencies of both the disease and a particular gene), for finding an association between a particular gene and the disease.218 Hence, an association may be found where no true causal relation exists.

Animal model genetic studies

There are several approaches to assessing the effects of particular gene knockouts in mouse models of schizophrenia. The first begins from the genetic manipulation approach. In this approach, specific genes of interest are mutated (e.g., loss of functionality, referred to as a gene knockout; reduced functionality, referred to as a gene knockdown; increased functionality; or new functionality), such that any observed schizophrenia impairments can be attributed to the gene product’s effect.227 The second approach uses behavioural assessment of the offspring of mice that have undergone random mutagenesis, where random point mutations are induced throughout the genome. Once a behavioural impairment of interest has been identified, genetic analyses can localize the gene responsible for the impairment.227

Mouse models can be derived from a variety of different hypotheses regarding the etiology of schizophrenia. Some knockout models have focused genes that are implicated in neurotransmission, and others on specific schizophrenia risk factors that have been identified in human patients. Several specific genes have been implicated in schizophrenia on the basis of human linkage and association studies. These genes have been selected for discussion based on the strength of their association with schizophrenia, as reported in the literature.

Proline dehydrogenase (PRODH)

Proline dehydrogenase is a gene encoding the mitochondrial enzyme proline oxidase, which is involved in the metabolism of L-proline and in the transfer of redox potential across the mitochondrial membrane.228 This gene may also have an indirect influence on glutamate-mediated transmission, which was previously mentioned to be disrupted in schizophrenia.228 This gene is located on chromosome 22q11, an area that has been implicated in several linkage studies; if deleted, this dramatically increases the risk of schizophrenia.225 Studies have shown that polymorphisms in PRODH are associated with a reduction in bilateral frontal white matter density, decreased striatal volume and increased striatal–frontal functional connectivity, which are all consistent with neuroanatomic findings of schizophrenia.229,230

Several human association studies have implicated polymorphisms in PRODH with the development of schizophrenia. The first study to implicate PRODH examined a 1.5-Mb region of chromosome 22q11 to address the individual role of the genes in this area and their contribution to the development of schizophrenia.231 After PRODH had been implicated in this initial study, the authors sought to identify further mutations in this gene. They found that in children with early-onset schizophrenia there were missense mutations in the PRODH gene that may have lead to a reduced functionality of the gene.231 A missense mutation is a point mutation (i.e., alteration of a single nucleotide) resulting in the substitution of one amino acid by another during translation. The resulting protein can be nonfunctional, isomorphic, neomorphic, hypomorphic or hypermorphic. Another study implicated nonsense or missense mutations in PRODH in a subset of patients with schizophrenia, along with a condition of hyperprolinemia.232 A nonsense mutation is a point mutation resulting in the replacement of an amino-acid coding codon by a stop codon, thereby terminating the translation of the protein prematurely. The resulting protein is therefore truncated and often nonfunctional. Other studies have also implicated this gene in the development of schizophrenia in a Chinese family233 and in individuals affected with schizoaffective disorder.234 However, other studies have not found such an association.235237 Despite some inconsistency in the results, taken together, the evidence seems to suggest that mutations in PRODH are implicated in the etiology of schizophrenia.

Mutations in PRODH have also been associated with the development of particular schizophrenia-related endophenotypes in humans. An at-risk polymorphism in PRODH has been shown to attenuate PPI in healthy controls, as well as reduce performance verbal working memory tasks.238

In animal models, this gene, when overexpressed, also appears to modulate PPI in mice, but has no effect on locomotor activity.239 Proline dehydrogenase knockout mice show highly increased levels of proline in the plasma, decreased levels of glutamate, aspartate and GABA in the hypothalamus, and reduced levels of GABA and aspartate in the frontal cortex.240 These mice have been shown to have deficits in PPI and non-spatial hippocampus-dependent learning and memory, but are not hyperactive and do not display stereotyped behaviour or deficits in spatial working memory or have any particular gross morphologic brain abnormalities.240,241 In sum, these animal models replicate similar endophenotypic findings as the human studies and suggest the possibility that several additional biomarkers may be linked to this gene.

Catechol-O-methyltransferase (COMT)

Catechol-O-methyltransferase is another gene located on chromosome 22q11. It encodes a protein that degrades catecholamines, such as DA, in the neuronal synapses.242 In fact, patients with schizophrenia have increased expression of COMT glial cells in the PFC, possibly disrupting dopaminergic transmission in that area.243 Several studies have implicated mutations of this gene in the development of schizophrenia. It has been suggested that a Val/Met substitution at codon 108 and/or 158 increases the activity of the enzyme by destabilizing the active site structure of the enzyme, thereby reducing dopaminergic transmission in the dorsolateral PFC.244 Polymorphisms in COMT have also been associated with an increase in the grey matter of the right superior temporal gyrus and decreases in the volume of the hippocampus and parahippocampal gyrus.230,245

Though a promising candidate gene, the results of COMT human association studies are inconsistent. Several studies using different populations, as well as several meta-analyses, have shown that polymorphisms in COMT were highly associated with the development of schizophrenia, although not necessarily the Val/Met substitution, which may have no effect or only a modest effect, or may be itself highly associated with another polymorphism that has a causative effect.246250 Whereas the above-mentioned studies found associations between polymorphisms of COMT and schizophrenia, it is noteworthy that the studies are rather inconsistent in their results. Each association study generally investigates more than 1 polymorphism and reports significant findings on 1 or more of these, but any given polymorphism is not necessarily associated across studies. Nevertheless, despite these inconsistent results, COMT appears to be an important gene in the development of schizophrenia, although the Val/Met polymorphism may not confer as much risk as originally believed.242

Mutations in COMT have also been associated with several schizophrenia endophenotypes. Consistent with increased dopaminergic function in individuals with the Val/Met polymorphism in COMT, healthy volunteers and patients with schizophrenia show increased PPI.73,251 Furthermore, this polymorphism has frequently been associated with WSCT performance, although recent studies and a meta-analysis have found that the association may only be small, although significant.252,253 Finally, mutations in COMT have also been associated with social anhedonia.254

In animal model studies, male homozygous Comt knockout mice showed a 2- to 3-fold increase in DA levels, although female mice did not.255,256 These knockout mice also exhibit decreased social behaviour, consistent with human studies, but no hyperactivity, stereotypy (i.e., perseverative behaviour) or deficits in PPI.256 The lack of a disruption in the latter 2 biomarkers appears to be inconsistent with the aforementioned human studies.

Neuregulin-1 (NRG1)

Neuregulin-1 is a gene located on chromosome 8p that has a role in the expression and activation of glutamate, GABA and other neurotransmitters, as well as additional roles in neurodevelopment, specifically cellular differentiation and neuronal development and migration.257 A variant in the NRG1 promoter region has been associated with decreased frontal and temporal lobe activation, the development of psychotic symptoms, cognitive deficits and increased volume of the lateral ventricles.258,259

Human association studies have suggested that mutations in NRG1 are highly implicated in the development of schizophrenia. The main haplotypes that have been associated with schizophrenia have been localized to both the 3′ and 5′ ends of the gene.260 Studies of various races have shown a significant association between schizophrenia, as well as psychotic features related to other disorders, and haplotypes of NRG1.261263 However, some studies have failed to find an association between mutations in NRG1 and schizophrenia.264,265 A recent meta-analysis of 13 human association studies of NRG1 found that 6 polymorphisms in NRG1, as well as at-risk haplotypes, showed a strong and consistent positive association with schizophrenia.266 Another recent meta-analysis also found several polymorphisms of NRG1 that were significantly associated with the development of schizophrenia.267 Interestingly, some haplotypes have been associated with the deficit form of schizophrenia, whereas others have been associated with the nondeficit form of schizophrenia.261 Again, as with the results of COMT, the various haplotypes associated with schizophrenia are not always consistent across studies, and no endophenotypes have been associated with this gene as of yet. Nevertheless, taken together, the results suggest a significant role for NRG1 in the development of schizophrenia.

Although results have been inconsistent across studies, in animal model studies Nrg1 hypomorphic mice have shown modest baseline locomotor hyperactivity, but normal amphetamine-, PCP– and MK-801–induced hyperactivity. They also showed PPI and LI disruptions, as well as increased drug-dependent disruption of PPI, but no spatial memory deficits.263,268,269 No differences were reported in terms of DA or serotonin receptor numbers.269 These mouse models are consistent with the literature on schizophrenia in humans and elucidate some possible schizophrenia-related biomarkers that can now be specifically examined in human patients.

Disrupted in schizophrenia 1 (DISC1)

Disrupted in schizophrenia 1 was originally discovered from a balanced translocation between chromosome 1 and 11, in which parts of chromosome 1 and 11 were interchanged.242,270 The 1q breakpoint was found to involve 2 genes: DISC1 and DISC2; the latter is thought to be a non–protein coding regulatory gene.270 Given the minimal number of genes at the breakpoint location on chromosome 11, it is not thought that any genes on this chromosome confer risk for the development of schizophrenia.270 The protein DISC1 appears to interact with a number of proteins that are important in neurite growth and neuronal migration and may also impact granule cell migration in the dentate gyrus.271,272 The neurodevelopmental roles of DISC1 render it of potential importance for understanding hypothesized brain development abnormalities in schizophrenia. Studies have suggested that polymorphisms in DISC1 may lead to a volumetric reduction of the supramarginal gyrus, a part of the posterior parietal cortex, a reduction in grey matter in the PFC, specifically in the superior frontal gyrus and anterior cingulate gyrus, impairments in PFC function and increased severity of positive symptoms.273275

The DISC1 translocation was first identified in a Scottish family and was associated with schizophrenia, bipolar disorder, major depression and other psychiatric disorders.270,276 These results may indicate that disruptions in DISC1 may not be unique to schizophrenia, but they may nonetheless confer risk for the development of the disorder. Polymorphisms of DISC1 have also been associated with schizophrenia and schizoaffective disorder in multiple populations of various races.277279 A recent meta-analysis and association study found a significant association between polymorphisms in DISC1 and schizophrenia in a European sample, and the meta-analysis revealed evidence for a common risk interval extending from intron 4 to 6, although several other polymorphisms were found to be significantly associated with schizophrenia across studies.280 Interestingly, it has been reported that the genes showing the strongest association were in the area of the breakpoint previously described.270,277 Polymorphisms of the DISC1 gene have only been associated with 1 endophenotype: social anhedonia.281

In animal model studies, working memory has been shown to be disrupted in a mouse strain that contains a natural mutation in Disc1.98 Behaviourally, studies have shown that mutations in Disc1 result in impairments in PPI, LI, social interaction and impairments in spatial and working memory.282285 A 22-hour locomotor test has also suggested that Disc1 knockout mice are hyperactive in the dark phase of the light-dark cycle.284 Disrupted in schizophrenia 1 knockout mice have also been shown to have enlarged lateral ventricles, reduced cerebral cortical volume, reduced neuronal proliferation, reduced GABAergic interneurons in the hippocampus and cortex and attenuated neurite outgrowth in primary cortical neurons.284,285 The behavioural and brain abnormalities are all reminiscent of those found in patients with schizophrenia. Therefore, these results suggest that mutations in DISC1 have a significant role in the development of schizophrenia-related biomarkers.

D-amino acid oxidase activator (DAOA)

D-amino acid oxidase activator, also known as G72, which is located on chromosome 13q, has also been associated with the development of schizophrenia.242 The protein, DAOA, is thought to be indirectly involved in glutamatergic transmission.242 Given that DAOA interacts with a protein called D-amino acid oxidase (DAO), which metabolizes a modulator of NMDA receptors (i.e., receptors for the glutamatergic system), it is thought that DAOA is involved in schizophrenia by influencing these receptors.286

Several studies have found significant associations between polymorphisms in DAOA and multiple racial groups.286288 Other studies have been more inconsistent in their results. One study was unable to replicate associations for previously reported polymorphisms but discovered positive associations with 2 new polymorphisms, which the authors potentially attributed to chance.289 A recently conducted meta-analysis found significant associations between polymorphisms in DAOA and schizophrenia in Chinese populations, but the associations lost significance when Korean populations were incorporated into the analysis.290 Other studies have altogether failed to find an association between polymorphisms in DAOA and schizophrenia.264,265,291 A large meta-analysis of 49 studies found a weak association between schizophrenia and mutations in DAOA, and there was a great deal of heterogeneity among the associated risk alleles.292 Taken together, the results suggest that DAOA may play only a minor role in the development of schizophrenia.

It has also been suggested that mutations in DAOA may also be involved in modulating the endophenotype of working memory. Interestingly, healthy individuals who carried the high-risk allele of DAOA had better memory performance than those who did not.293

From an animal model perspective, there has been little work done to characterize the behavioural abnormalities resulting from disruptions of this gene. In the human genome, there are 2 overlapping genes, G72 (DAOA) and G30, the former of which is only expressed in primates.294 Therefore, the human G72/G30 genomic region was spliced into mouse DNA to create the model. This transgenic mouse showed impairments in PPI, uncoordinated locomotor activity and increased sensitivity to PCP administration.294 It is unclear how and why behavioural deficits would appear in these mice if the spliced gene region did not contain schizophrenia-related polymorphisms. Nevertheless, this study suggests that the DAOA gene may be important in the development of schizophrenia-related biomarkers.

Dystrobrevin binding protein I (DTNBP1)

Dystobrevin binding protein I, also known as dysbindin, is located on chromosome 6p and is thought to influence glutamatergic transmission, although its exact function is not well understood.242 It has been suggested that reduced expression of DTNBP1 mRNA in the PFC and the glutamatergic terminals of the hippocampal formation are associated with schizophrenia, although the exact mechanisms are not well understood.295,296

A strong association has been found between polymorphisms in DTNBP1 and schizophrenia in multiple racial cohorts, some particularly at the 3′ end of the gene.265,297299 However, several studies have failed to confirm this association.300302 Again, as with many of the previously discussed genes, there were multiple inconsistencies in and failures to replicate the positively associated polymorphisms across studies. A recent meta-analysis of 12 studies suggested that there is only a weak association of 1 single nucleotide polymorphism (SNP) in DTNBP1 and schizophrenia and that associations of additional SNPs with the disorder may be the direct result of only 1 study.303 These results suggest that some form of association may exist between mutations in DTNBP1 and schizophrenia, although the exact extent to which mutations of DTNBP1 are involved in the development of schizophrenia is still unclear.

In terms of endophenotypes, DTNBP1 seems to be involved in memory performance. Patients with an at-risk haplotype of DTNBP1 performed more poorly than patients who were noncarriers on a spatial memory task.304 Similarly, a protective haplotype of DTNBP1 was found to increase memory performance in healthy controls, although it had no performance effect in patients with schizophrenia.305

Recently, there has been a proliferation of animal model studies that have examined Dtnbp1 in the sandy mouse, which does not express the Dtnbp1 gene.306 Studies have shown that Dtnbp1 knockout mice display reduced social interaction, impairments in long-term memory retention, spatial memory and working memory, as well as hyperactivity.306309 These mice also have a reduction in the steady state levels of synapsin in the hippocampal formation, a reduction in DA, but normal glutamate levels.308,309 Given the findings from both animal model and human association studies, it would appear that DTNBP1 is an influential gene in the development of schizophrenia-related biomarkers and schizophrenia, respectively.

Reticulon-4 and the reticulon-4 receptor (RTN4 and RTN4R)

Reticulon-4 is also located on chromosome 22q11 and is a glycosylphosphatidilinositol-linked protein, which contains multiple leucine-rich repeats that bind to the Nogo-66 protein.310 Nogo-66 is a myelin-associated protein that inhibits the outgrowth of neurites and nerve terminals.310 In addition, 2 other proteins thought to be involved in the inhibitory components of myelin also bind to the reticulon-4 receptor.310 Investigations of these genes are currently in their infancy. Several studies have shown that polymorphisms in the Nogo-66 gene itself (RTN4), located on chromosome 2p, particularly a CAA insert in the 3′ region of the gene, confers risk for the development of schizophrenia.311,312 However, 3 other studies failed to replicate this result.313,314

When RTN4R has been examined, the results have been mixed. Several studies have found associations between polymorphisms of RTN4R and schizophrenia in multiple populations, ranging in strength of positive associations.310,315,316 One of these studies also found that some mutant alleles resulting from missense mutations were found in patients who were strongly resistant to neuroleptic treatment.310 However, another study did not find such an association in a Chinese population.317 Although RTN4R is a positional candidate gene, it has not been associated with any particular endophenotypes, and, therefore, further work on this gene is necessary to determine more conclusively whether it is a risk factor for the development of schizophrenia.

There are few studies that have assessed behavioural abnormalities in Rtn4r knockout mice. One study has shown that Rtn4r knockout mice have locomotor deficits but no impairments in PPI or working memory.316 However, another study has suggested that Rtn4r knockout mice have impairments in spatial working memory.315 These results seem promising, but given the inconsistent results of human association and animal model studies, the involvement of RTN4R in the development of schizophrenia-related biomarkers and schizophrenia is still in question.

The regulator of G-protein signalling 4 (RGS4)

The regulator of G-protein signalling 4 is a gene located on chromosome 1q and mediates postsynaptic transduction in dopaminergic, glutamatergic and serotonergic signalling pathways.318 One study has suggested that allelic variations in RGS4 are associated with changes in the functional pathways involved in working memory, grey matter structural connectivity and white matter volume.318

Several studies, including a meta-analysis, have reported associations between polymorphisms in RGS4 and schizophrenia.319321 However, other studies, including a subsequent meta-analysis, have found no such associations.265,322,323 Whereas RGS4 is a positional and functional candidate for the development of schizophrenia, it has not been associated with specific endophenotypes, and the results, when taken together, render it difficult to determine the extent to which RGS4 is a risk factor for the development of schizophrenia.

In terms of animal model research, 1 study has reported that Rgs4 knockout mice did not show impairments in PPI or working memory.324 Given these findings, as well as the inconsistency in the human literature, RGS4 may not be as promising a candidate gene as first expected.

Zinc finger DHHC-type containing 8 (ZDHHC8)

Zinc finger DHHC-type containing 8 is located on chromosome 22q11 and is a putative palmitoyltransferase protein expressed in the brain, particularly in the cortex and the hippocampus, that may be involved in synaptic transmission and post-translational modification of neurotransmitter systems.325

An initial scan of the 22q11 locus revealed 3 polymorphisms in ZDHHC8 that were associated with schizophrenia.326 Subsequent studies have confirmed positive associations between mutations in ZDHHC8 and schizophrenia across a number of racial groups, as well as a sex-related heterogeneity of allele transmission, whereby female patients showed a stronger association between mutations in ZDHHC8 and schizophrenia.325,327 However, multiple studies have been unable to replicate these findings across various racial groups, although modest evidence of a sex-related heterogeneity of allele transmission has been confirmed.328,329 Whereas ZDHHC8 is a positional candidate gene, no endophenotypes have been associated with it and, given the inconsistent results, it is difficult to make a conclusive determination about its role in the development of schizophrenia.

In animal model research, a mouse model where both Comt and Zdhhc8 were overexpressed showed decreased locomotor activity but no differences in PPI.239 Given that Comt knockout mice have shown locomotor impairments, one might hypothesize that Zdhhc8, or the interaction of the 2 genes, is responsible for the modulation of locomotor activity in these mice. A Zdhhc8 knockout mouse has shown deficits in PPI, decreased locomotor activity and decreased sensitivity to MK-801 administration, but no gross morphologic brain abnormalities.325 It has also been demonstrated that these mice have a decreased density of dendritic spines and glutamatergic dendrites, as well as impairments in dendritic growth.330

Conclusions of genetic animal model research

Taken together, the results from these animal models suggest that mutations in PRODH, NRG1, DISC1 and DTNBP1 may confer greater risk for the development of schizophrenia, as they have been shown to be causally related to many schizophrenia-related abnormalities. It is important to note that these are the same genes that tend to be consistently associated with the development of schizophrenia in humans. In addition to these genes, mutations in COMT have strong human but not animal model support. Conversely, animal models with weaker schizophrenia-related endophenotypic presentation tend to involve mutations in genes (i.e., DAOA, RTN4R, RGS4 and ZDHHC8) that are not as consistently associated with the development of schizophrenia. This may suggest several possible conclusions. First, these genes may only be involved in conferring a small degree of risk, resulting in the presence of fewer schizophrenia-related abnormalities in animal studies and greater inconsistency in human association studies. Second, the gene or the specific gene mutation may not be involved in the development of schizophrenia, and the presence of biomarkers and associations may be artifactual. Third, some of these genes may require an interaction with other risk genes to confer risk. For example, RGS4, DAOA and DISC1 all show evidence of epistatis with COMT, suggesting that there may be an increased risk for schizophrenia based on an interaction between the genes.331 Single gene knockout animal models would be unable to detect these interactions.

Genetic knockout models of schizophrenia, particularly those based on genes that are known risk factors, are useful in that they can show strong etiologic relevance to schizophrenia. These studies enable a determination of the causal link between the disruption of specific genes and schizophrenia-related biomarkers independent of other genetic or neurodevelopmental risk factors. In addition, they also permit an assessment of underlying mechanisms and downstream effects of the gene, including how these effects relate to the presence of biomarkers. For example, studies on the Dat knockout mouse have shown that both D1 and D2 receptors may underlie the expression of hyperactivity, whereas D1 may regulate stereotypy, and D2 may regulate sensorimotor gating.156 Interestingly, results show that not all schizophrenia-related abnormalities are observed in a particular knockout model, which suggests that the disruption of a particular gene may modulate some, but not all, endophenotypes. Genetic animal models, therefore, can help determine which genes may modulate particular biomarkers and endophenotypes, but by no means approximate the complexity of the etiology of schizophrenia.

The complementary nature of human and animal genetic studies

Human genetic studies are indispensible in understanding the genetic basis of schizophrenia. Linkage studies are necessary to determine chromosomal locations of risk genes, and association studies are important to determine the relation between specific genes and schizophrenia. However, human association studies only provide correlational results and have typically yielded inconsistent results. These inconsistencies may result from a previously associated gene that may not be associated in a new study. Additionally, studies that examine different polymorphisms of a given gene may find, quite correctly, different results. Although some human association studies do not show an association between these genes and the development of schizophrenia, the results of animal models have suggested that there may be causal relations in some cases. Hence, null results obtained in human association studies could be explained by alternative reasons, rather than the lack of an association.

The failure to obtain an association does not mean that one does not exist in some human populations. The polygenic multifactorial threshold model suggests that there are a greater number of risk-increasing polygenes than the number of genes necessary to cross the threshold.44 Hence, in one family or population, genes A, B and C may increase the risk for schizophrenia, whereas in another population genes B, C and D may increase the risk. A negative association for gene A in the second population does not mean that gene A is not a risk factor for schizophrenia. Rather, it may mean that gene A is not a risk factor in that particular population. In fact, it was noted in a previous meta-analysis of DAOA that a significant association in a Chinese population was rendered nonsignificant after the addition of a Korean population.290 This exemplifies the fact that the failure to find an association may be population-specific and that a particular gene may be associated in some populations but not others. Regardless, the gene would still likely be an important risk factor.

Another reason for inconsistent results is the heterogeneity of the polymorphisms (i.e., SNPs) that are examined. It can easily happen that one study finds an association with a particular polymorphism, but another study does not find an association with a different polymorphism in the same gene. This does not imply that the gene itself is not a risk factor but rather that a particular polymorphism of that gene is not a risk factor. Hence, the failure to replicate a positive association might be the result of using an SNP that was designed to find a polymorphism in a different part of the gene.

Finally, there may be statistical reasons why an association may not be found. Many recent studies have assessed multiple SNPs simultaneously. Often, results are presented individually for each polymorphism, where 1 or more SNPs may be significant. However, owing to the large number of SNPs simultaneously assessed, the possibility exists that some of the significant associations occurred by chance. Hence, a statistical correction must be made to the α level, rendering it more difficult to find significant associations. Only associations that are highly significant remain so after the correction, whereas those that are less significant become nonsignificant.

Clearly, there are a number of methodologic issues inherent in genetic research in humans with schizophrenia. It is in this domain where the contribution of animal models is most evident. Human association studies provide the means for identifying relevant risk genes, but these studies have their limitations. By experimentally manipulating a gene in a mouse, one can ensure that all animals in the experimental group contain the particular risk gene with a particular polymorphism. This avoids the problem of examining populations that may not have polymorphisms in a particular gene of interest. Furthermore, environmental and developmental factors can be controlled across the lifespan of the mouse, thereby removing such factors as experimental confounds. Animal models avoid the correlational nature of human linkage and association studies by using an experimental paradigm. Therefore, the effect of a particular gene can be causally related to the development of a particular endophenotype. Multiple lines of knockout mice can also be used, such that different lines each have a different mutation in the same gene, thereby permitting a controlled and more thorough investigation of the particular regions of the gene that might confer risk.

However, genetic animal models are not without their limitations. Generally, animal models have involved the study of 1 gene at a time; this is a rather artificial way to study schizophrenia, given the polygenic nature of the disorder. The reliance on the use of endophenotypes, while beneficial in some ways, is also a limitation, as the same endophenotype can be found across a number of disorders. Animal models of autism, for example, show deficits in social interaction.131 Therefore, if a particular gene knockout mouse only shows impairments in 1 biomarker, it is difficult to conclude that that the model is one of schizophrenia as opposed to another psychiatric disorder for which the biomarker might be relevant. It is, therefore, the complementary nature of the human and animal literature that can provide the clearest picture. A causal link between a particular gene and a schizophrenia-related biomarker in an animal model is most appropriate in the context of a plausible association between that gene and schizophrenia in humans.

Evaluation of animal models of schizophrenia

The previous sections have discussed schizophrenia-related abnormalities that are associated with many different kinds of animal models, including pharmacologic, neurodevelopmental and genetic animal models. The etiologic validity of these models depends largely on how closely they can approximate known schizophrenia risk factors. The resulting abnormalities can then be causally linked with each specific experimental manipulation. As we have noted throughout this review, many of these abnormalities are present in other psychiatric disorders, and the presence of multiple schizophrenia-related abnormalities does not necessarily imply that there is a relation to the development of schizophrenia as opposed to other disorders. However, this relation becomes more valid in the context of more etiologically valid models.

In such models, one can assess the impact of each individual manipulation, whether it is a genetic knockout or a replication of an obstetric complication, on behavioural, morphologic and neurochemical levels among others. Given the heterogeneity of schizophrenia symptom presentation, this understanding is crucial in delineating possible subsets of the population that may differ in terms of phenomenology or treatment outcome depending on the presence of certain risk factors and their associated biomarkers. Interestingly, many of the aforementioned animal models seem to share similar, although sometimes distinct, schizophrenia-related abnormalities. For example, deficits in PPI, LI and working memory are among some of the more consistent behavioural changes that are observed in these models. It is noteworthy that many of these animal models have distinct etiologic origin. This may suggest a number of possibilities.

First, these animal models may highlight the multifactorial nature of the disorder, such that presence of multiple risk factors is required for schizophrenia to develop.44 From a behavioural perspective, each gene should result in particular abnormalities, such that the additive effect would lead to the development of schizophrenia symptoms.55 Thus, one could hypothesize that a combination of risk factors would result in greater impairments. Some biomarkers are common among risk factors, such that the combination of the risk factors could result in a more severe impairment. Alternatively, some biomarkers are distinctive to each risk factor, such that the combination of risk factors could result in a greater array of abnormalities. We discuss the necessity of multiple risk factor models in the subsequent section. However, based on the convergence and divergence of schizophrenia-related endophenotypes in etiologically distinct animal models, this may indicate that as risk factors accumulate, the severity and number of abnormalities should increase, thereby leading to the development of actual schizophrenia symptomatology.

A second possibility is that many of these risk factors could combine in a neural common pathway that ultimately leads to the development of schizophrenia. Some authors have hypothesized that many of the risk factors culminate in a disruption of neurotransmitter function, thereby affecting multiple brain areas and leading to the development of schizophrenia.153,154 Such hypotheses would be difficult to test in animal models, as they cannot model the full complexity of the etiology or phenomenology of schizophrenia. Animal models can never provide a surrogate for schizophrenia as they can only model the effects of 1 or potentially a few risk factors at a time. Unfortunately, animal models cannot model complex genetic and neurodevelopmental interactions or epigenetic effects, which may be crucial in the development of the disorder. In fact, this limitation may underlie some of the discrepancies between human and animal studies of a particular risk factor. For example, our earlier discussion indicated that COMT had more support as a schizophrenia risk factor in human studies than in animal model studies. Conversely, the discussion indicated that DTNBP1 appeared to have more support in the animal model studies than in human association studies. Although this may be owing to the lack of homology between human and rodent neurobiology, it may also reflect both protective and nonprotective interactions that are not modelled by single-gene knockout animal models.

The main use of animal models of schizophrenia risk factors is that they can permit an understanding of the mechanism of each risk factor in isolation. However, to test the aforementioned hypotheses, animal models that replicate multiple risk factors must be created. However, such models are currently rare, despite being crucial for schizophrenia research (see the following section). Most research has focused on the level of behavioural, morphologic or neurotransmitter abnormality, but an in-depth understanding of the mechanism of action of each risk factor would be important. This becomes particularly evident in the area of treatment. As many of the animal model studies have highlighted, schizophrenia-related biomarkers have been associated with a number of different etiologies, including neurotransmitter disruption, abnormalities in brain circuitry, neurogenesis disruption and genetic effects that lead to both neurodevelopmental and neurotransmitter disruptions. This suggests that the etiology of schizophrenia is diverse and complex, and this notion can impact treatment. When one examines the etiology of a variety of animal models, current dopaminergic treatment options for schizophrenia may be largely incomplete and ineffective at targeting the underlying pathology of schizophrenia.332 This suggests that more disease-centred treatments, as opposed to drug-centred treatments, may be required to successfully treat schizophrenia.332 For example, it has been suggested that novel treatments involving viral-vector gene transfer into the brain could be used to ameliorate psychotic symptoms or modify the course of the illness.333 Animal models are not only well-suited to understand the mechanisms of schizophrenia, but could be particularly useful in testing the effects of more specialized treatment.

Animal models are clearly important in understanding the mechanisms underlying particular risk factors and related biomarkers. However, animal model research must still take place to more effectively complement research on schizophrenia in humans, and allow for a further unveiling of the etiology and pathogenesis of the disorder.

Future directions for animal models of schizophrenia

Animal models are useful in complementing the understanding of research on schizophrenia in humans, although they are not without disadvantages. Several important areas for further improvement have been highlighted in this review.

First, as previously discussed, background strain effects can alter the expression of certain behavioural abnormalities. Certain background strains may inadvertently contain mutations that can render the findings of a study ineffectual. For example, previous research has determined that mutations in Prodh resulted in deficits in PPI.240 However, more recent research showed that the background strain used in this study had a mutation in another schizophrenia risk gene (i.e., Disc1) that led to PPI disruptions.334 Therefore, there is a genetic confound making it unclear whether it is mutation of the Disc1 gene or mutation of both the Prodh and Disc1 genes that resulted in PPI deficits. In addition, there is a large accumulation of evidence indicating that different mouse strains display different baseline behaviours.132,335 Similarly, some strains may have sensory impairments. For example, several strains contain mutations that cause retinal degeneration or impaired vision, rendering them unsuitable in a behavioural assay that involves visual information.336,337 To remove genetic, sensory or behavioural confounds, and to ensure consistency in endophenotype expression and comparability across studies, a standard background strain, such as C57Bl/6J, should be used in mouse models of schizophrenia.

Second, the results of human association studies are often inconsistent owing to the diversity of polymorphisms studied. Similarly, this issue can also apply to genetic knockout research. The general model of knockout research is to inactivate the gene, based on the supposition that such an inactivation would have the largest effect on behaviour. However, such drastic genetic manipulations are not necessarily etiologically valid, as many human risk genes are associated with polymorphisms rather than complete gene deactivation. Similarly, different mutations within the same gene have been shown to lead to different behavioural profiles.282 This suggests that small and distinct changes in the gene, which could result in altered protein folding, may cause different behavioural effects. Therefore, it is important to create animal models that reflect the polymorphisms or physiologic effects of identified risk genes. Thus, although it is important to determine the general effect of the gene on the development of schizophrenia by deactivating it, it is also important to target specific polymorphisms identified in the literature on schizophrenia in humans.

Another weakness of current genetic models is that they often only target 1 gene at a time. Although this permits an understanding of the gene’s effects without additional confounds, the polygenic nature of the disorder is not modelled. This review has illustrated that mutations in different genes can lead to distinct or similar biomarkers, and the possible combination of these genes could lead to a greater number or degree of behavioural impairments. To assess the contribution of multiple polygenes, as is suspected in schizophrenia in humans, it would be important to create multiple-gene knockout mice. Although this would be a complex and laborious undertaking, it would allow for an analysis of the relative contribution, as well as the interaction and additive effects, of each gene to the development of schizophrenia. It would certainly be interesting to knock out a combination of genes of relatively strong effect, such as Prodh, Nrg1, Disc1 and Dtnbp1, to determine the number or severity of the resulting abnormalities. It would also be interesting to combine knockouts of the aforementioned genes with those of more modest effect, such as Daoa, Rtn4r, Rgs4 and Zdhhc8, to detect interactions that might exist between these different classes of genes. Another avenue for modelling the multifactorial nature of the disorder would be to combine genetic knockout with environmental risk models. As previously noted, obstetric complications take on relevance to the development of schizophrenia within the context of a genetic predisposition.179 Therefore, single or multiple knockout mice could be combined with single or multiple environmental risk factors. These animal models would permit a more complete understanding of the contribution and interaction of all possible risk factors and would, therefore, more completely model the etiology of the disorder.

Another important factor is the use of a battery of behavioural tests when assessing the effect of a particular gene on the development of schizophrenia. This provides a more complete behavioural characterization and permits a comparison between the behavioural abnormalities observed across models. Since a specific abnormality can be observed across multiple psychiatric disorders, it is important to obtain a convergence of schizophrenia-related abnormalities to determine that a particular gene is involved in the development of the disorder. Certainly, other tests could be added to many of the aforementioned studies to assess for additional schizophrenia-related behaviours, and provide a more complete behavioural characterization. Therefore, the greater the number of tests used, the greater the amount of convergence, and the greater the validity of the animal model. It may also be important to assess for behaviours that are not related to schizophrenia (i.e., depression and anxiety) to determine whether certain abnormalities are modulated through the pathways of other psychiatric disorders and whether there is a common genetic basis between different disorders.

Conclusion

Research on schizophrenia in humans is indispensible in understanding the relation between risk factors and the development of the disorder. However, taken by itself, research in human patients has limitations. Animal models of schizophrenia can complement research in humans by employing experimental paradigms, leading to possible conclusions regarding causation, which cannot be obtained from correlational data. In addition, such models can eliminate many confounds expected in schizophrenia research in humans by studying only 1 risk factor at a time. Such studies can, therefore, elucidate the etiology of schizophrenia by examining the causal pathways involved in the development of the disorder. Nevertheless, animal models cannot provide a complete picture of schizophrenia. One cannot obtain the complete clinical picture in a rodent owing to the lack of symptom presentation and the less than perfect homology in the CNS of both species. Hence, extrapolations to schizophrenia, on the basis of findings from animal model research, must be done cautiously.

In general, animal model research, although important and often well founded in the literature, is still incomplete. Many animal models have not been fully characterized on a phenotypic level using a large battery of tests and similar background strains. In addition, research using different knockout models of the same gene, resulting from mutations in different locations, is still in its infancy. Perhaps more importantly, multiple knockout and gene–environment interaction models are lacking and must be developed. Given the polygenic nature of the disorder, such models would be useful in better approximating the etiology of schizophrenia.

Nevertheless, the convergent evidence from human and animal studies can further elucidate the etiology, the mechanisms and the phenotype of the disorder. In addition, the complementary nature of these 2 types of studies can help circumscribe the boundaries of knowledge in each case. By applying the knowledge gained in both human and animal research on schizophrenia, advances in understanding the etiology, mechanisms and neurobiology of the disorder can be achieved and refinements in treatment procedures in both clinical and research settings can be implemented.

Acknowledgements

The authors thank Drs. K.P. Ossenkopp, S. MacDougall-Shackleton and E. Hayden for their advice and assistance during the preparation of this manuscript. This research was supported by grants from the National Sciences and Engineering Research Council of Canada (grant number A0669 to D.P.C; Canada Graduate Scholarship–doctoral student support to N.L.L.), and the Social Science and Humanities Research Council of Canada (grant number CID 28195 to R.W.J.N.).

Footnotes

Competing interests: See above.

Contributors: Mr. Lazar and Dr. Neufeld designed the review. Mr. Lazar acquired the data, which Dr. Cain analyzed. Mr. Lazar wrote the article. All authors reviewed the article and approved publication.

References

  • 1.American Psychiatric Association. Diagnostic and statistical manual of mental disorders. Washington (DC): American Psychiatric Association; 2000. [Google Scholar]
  • 2.Mao CV, Hori E, Maior RS, et al. A primate model of schizophrenia using chronic PCP treatment. Rev Neurosci. 2008;19:83–9. doi: 10.1515/revneuro.2008.19.2-3.83. [DOI] [PubMed] [Google Scholar]
  • 3.Miczek KA, Yoshimura H. Disruption of primate social behavior by d-amphetamine and cocaine: differential antagonism by antipsychotics. Psychopharmacology (Berl) 1982;76:163–71. doi: 10.1007/BF00435272. [DOI] [PubMed] [Google Scholar]
  • 4.Nicholson IR, Chapman JE, Neufeld RW. Variability in BPRS definitions of positive and negative symptoms. Schizophr Res. 1995;17:177–85. doi: 10.1016/0920-9964(94)00088-p. [DOI] [PubMed] [Google Scholar]
  • 5.Buchanan RW. Persistent negative symptoms in schizophrenia: an overview. Schizophr Bull. 2007;33:1013–22. doi: 10.1093/schbul/sbl057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Tamminga CA, Holcomb HH. Phenotype of schizophrenia: a review and formulation. Mol Psychiatry. 2005;10:27–39. doi: 10.1038/sj.mp.4001563. [DOI] [PubMed] [Google Scholar]
  • 7.Heinrichs RW. The primacy of cognition in schizophrenia. Am Psychol. 2005;60:229–42. doi: 10.1037/0003-066X.60.3.229. [DOI] [PubMed] [Google Scholar]
  • 8.Lieberman JA, Perkins D, Belger A, et al. The early stages of schizophrenia: speculations on pathogenesis, pathophysiology, and therapeutic approaches. Biol Psychiatry. 2001;50:884–97. doi: 10.1016/s0006-3223(01)01303-8. [DOI] [PubMed] [Google Scholar]
  • 9.Marenco S, Weinberger DR. The neurodevelopmental hypothesis of schizophrenia: following a trail of evidence from cradle to grave. Dev Psychopathol. 2000;12:501–27. doi: 10.1017/s0954579400003138. [DOI] [PubMed] [Google Scholar]
  • 10.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]
  • 11.Kessler RM, Woodward ND, Riccardi P, et al. Dopamine D2 receptor levels in striatum, thalamus, substantia nigra, limbic regions, and cortex in schizophrenic subjects. Biol Psychiatry. 2009;65:1024–31. doi: 10.1016/j.biopsych.2008.12.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lyon GJ, Abi-Dargham A, Moore H, et al. Presynaptic regulation of dopamine transmission in schizophrenia. Schizophr Bull. 2009 Jun 12; doi: 10.1093/schbul/sbp010. [Epub ahead of print] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Abi-Dargham A, Laruelle M. Mechanisms of action of second generation antipsychotic drugs in schizophrenia: insights from brain imaging studies. Eur Psychiatry. 2005;20:15–27. doi: 10.1016/j.eurpsy.2004.11.003. [DOI] [PubMed] [Google Scholar]
  • 14.Ananth J, Burgoyne KS, Gadasalli R, et al. How do the atypical antipsychotics work? J Psychiatry Neurosci. 2001;26:385–94. [PMC free article] [PubMed] [Google Scholar]
  • 15.Faustman WO, Bardgett M, Faull KF, et al. Cerebrospinal fluid glutamate inversely correlates with positive symptom severity in unmedicated male schizophrenic/schizoaffective patients. Biol Psychiatry. 1999;45:68–75. doi: 10.1016/s0006-3223(98)00207-8. [DOI] [PubMed] [Google Scholar]
  • 16.Moghaddam B. Targeting metabotropic glutamate receptors for treatment of the cognitive symptoms of schizophrenia. Psychopharmacology (Berl) 2004;174:39–44. doi: 10.1007/s00213-004-1792-z. [DOI] [PubMed] [Google Scholar]
  • 17.Tsai G, Passani LA, Slusher BS, et al. Abnormal excitatory neurotransmitter metabolism in schizophrenic brains. Arch Gen Psychiatry. 1995;52:829–36. doi: 10.1001/archpsyc.1995.03950220039008. [DOI] [PubMed] [Google Scholar]
  • 18.Javitt DC, Zukin SR. Recent advances in the phencyclidine model of schizophrenia. Am J Psychiatry. 1991;148:1301–8. doi: 10.1176/ajp.148.10.1301. [DOI] [PubMed] [Google Scholar]
  • 19.Akbarian S, Kim JJ, Potkin SG, et al. Gene expression for glutamic acid decarboxylase is reduced without loss of neurons in prefrontal cortex of schizophrenics. Arch Gen Psychiatry. 1995;52:258–66. doi: 10.1001/archpsyc.1995.03950160008002. [DOI] [PubMed] [Google Scholar]
  • 20.Volk D, Austin M, Pierri J, et al. GABA transporter-1 mRNA in the prefrontal cortex in schizophrenia: decreased expression in a subset of neurons. Am J Psychiatry. 2001;158:256–65. doi: 10.1176/appi.ajp.158.2.256. [DOI] [PubMed] [Google Scholar]
  • 21.Benes FM, Khan Y, Vincent SL, et al. Differences in the subregional and cellular distribution of GABAA receptor binding in the hippocampal formation of schizophrenic brain. Synapse. 1996;22:338–49. doi: 10.1002/(SICI)1098-2396(199604)22:4<338::AID-SYN5>3.0.CO;2-C. [DOI] [PubMed] [Google Scholar]
  • 22.Andreasen NC, Ehrhardt JC, Swayze VW, et al. Magnetic resonance imaging of the brain in schizophrenia. The pathophysiologic significance of structural abnormalities. Arch Gen Psychiatry. 1990;47:35–44. doi: 10.1001/archpsyc.1990.01810130037006. [DOI] [PubMed] [Google Scholar]
  • 23.Kelsoe JR, Jr, Cadet JL, Pickar D, et al. Quantitative neuroanatomy in schizophrenia. A controlled magnetic resonance imaging study. Arch Gen Psychiatry. 1988;45:533–41. doi: 10.1001/archpsyc.1988.01800300029003. [DOI] [PubMed] [Google Scholar]
  • 24.Schlaepfer TE, Harris GJ, Tien AY, et al. Decreased regional cortical gray matter volume in schizophrenia. Am J Psychiatry. 1994;151:842–8. doi: 10.1176/ajp.151.6.842. [DOI] [PubMed] [Google Scholar]
  • 25.Buchanan RW, Vladar K, Barta PE, et al. Structural evaluation of the prefrontal cortex in schizophrenia. Am J Psychiatry. 1998;155:1049–55. doi: 10.1176/ajp.155.8.1049. [DOI] [PubMed] [Google Scholar]
  • 26.Selemon LD, Goldman-Rakic PS. The reduced neuropil hypothesis: a circuit based model of schizophrenia. Biol Psychiatry. 1999;45:17–25. doi: 10.1016/s0006-3223(98)00281-9. [DOI] [PubMed] [Google Scholar]
  • 27.Harrison PJ. The neuropathology of schizophrenia. A critical review of the data and their interpretation. Brain. 1999;122(Pt 4):593–624. doi: 10.1093/brain/122.4.593. [DOI] [PubMed] [Google Scholar]
  • 28.Jakob H, Beckmann H. Prenatal developmental disturbances in the limbic allocortex in schizophrenics. J Neural Transm. 1986;65:303–26. doi: 10.1007/BF01249090. [DOI] [PubMed] [Google Scholar]
  • 29.Okugawa G, Sedvall GC, Agartz I. Reduced grey and white matter volumes in the temporal lobe of male patients with chronic schizophrenia. Eur Arch Psychiatry Clin Neurosci. 2002;252:120–3. doi: 10.1007/s00406-002-0370-9. [DOI] [PubMed] [Google Scholar]
  • 30.Sanfilipo M, Lafargue T, Rusinek H, et al. Volumetric measure of the frontal and temporal lobe regions in schizophrenia: relationship to negative symptoms. Arch Gen Psychiatry. 2000;57:471–80. doi: 10.1001/archpsyc.57.5.471. [DOI] [PubMed] [Google Scholar]
  • 31.Anderson JE, Wible CG, McCarley RW, et al. An MRI study of temporal lobe abnormalities and negative symptoms in chronic schizophrenia. Schizophr Res. 2002;58:123–34. doi: 10.1016/s0920-9964(01)00372-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Heckers S, Konradi C. Hippocampal neurons in schizophrenia. J Neural Transm. 2002;109:891–905. doi: 10.1007/s007020200073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Arnold SE. The medial temporal lobe in schizophrenia. J Neuropsychiatry Clin Neurosci. 1997;9:460–70. doi: 10.1176/jnp.9.3.460. [DOI] [PubMed] [Google Scholar]
  • 34.Bogerts B, Lieberman JA, Ashtari M, et al. Hippocampus-amygdala volumes and psychopathology in chronic schizophrenia. Biol Psychiatry. 1993;33:236–46. doi: 10.1016/0006-3223(93)90289-p. [DOI] [PubMed] [Google Scholar]
  • 35.Rajarethinam R, DeQuardo JR, Miedler J, et al. Hippocampus and amygdala in schizophrenia: assessment of the relationship of neuroanatomy to psychopathology. Psychiatry Res. 2001;108:79–87. doi: 10.1016/s0925-4927(01)00120-2. [DOI] [PubMed] [Google Scholar]
  • 36.Vita A, De PL, Silenzi C, et al. Brain morphology in first-episode schizophrenia: a meta-analysis of quantitative magnetic resonance imaging studies. Schizophr Res. 2006;82:75–88. doi: 10.1016/j.schres.2005.11.004. [DOI] [PubMed] [Google Scholar]
  • 37.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]
  • 38.Zaidel DW. Regional differentiation of neuron morphology in human left and right hippocampus: comparing normal to schizophrenia. Int J Psychophysiol. 1999;34:187–96. doi: 10.1016/s0167-8760(99)00076-8. [DOI] [PubMed] [Google Scholar]
  • 39.Csernansky JG, Schindler MK, Splinter NR, et al. Abnormalities of thalamic volume and shape in schizophrenia. Am J Psychiatry. 2004;161:896–902. doi: 10.1176/appi.ajp.161.5.896. [DOI] [PubMed] [Google Scholar]
  • 40.Keshavan MS, Rosenberg D, Sweeney JA, et al. Decreased caudate volume in neuroleptic-naive psychotic patients. Am J Psychiatry. 1998;155:774–8. doi: 10.1176/ajp.155.6.774. [DOI] [PubMed] [Google Scholar]
  • 41.Weinberger DR. Implications of normal brain development for the pathogenesis of schizophrenia. Arch Gen Psychiatry. 1987;44:660–9. doi: 10.1001/archpsyc.1987.01800190080012. [DOI] [PubMed] [Google Scholar]
  • 42.Cannon TD, Kaprio J, Lonnqvist J, et al. The genetic epidemiology of schizophrenia in a Finnish twin cohort. A population-based modeling study. Arch Gen Psychiatry. 1998;55:67–74. doi: 10.1001/archpsyc.55.1.67. [DOI] [PubMed] [Google Scholar]
  • 43.Murray RM, Castle DJ. Genetic and environmenal risk factors for schizophrenia. In: Gelder MG, Lopez-Ibor JJ Jr, Andreasen NC, editors. New Oxford textbook of psychiatry. Vol. 1. Toronto (ON): Oxford University Press; 2000. pp. 599–605. [Google Scholar]
  • 44.Pogue-Geile MF, Gottesman II. Schizophrenia: study of a genetically complex phenotype. In: Jones BC, Mormede P, editors. Neurobehavioral genetics: methods and applications. New York (NY): CRC Press; 1999. pp. 247–64. [Google Scholar]
  • 45.Karkowski L. Human family studies. In: Jones BC, Mormede P, editors. Neurobehavioral genetics: methods and applications. New York (NY): CRC Press; 1999. pp. 151–62. [Google Scholar]
  • 46.Carlier M, Spitz E. The twin method. In: Jones BC, Mormede P, editors. Neurobehavioral genetics: methods and applications. New York (NY): CRC Press; 1999. pp. 141–50. [Google Scholar]
  • 47.Gottesman II, Bertelsen A. Confirming unexpressed genotypes for schizophrenia. Risks in the offspring of Fischer’s Danish identical and fraternal discordant twins. Arch Gen Psychiatry. 1989;46:867–72. doi: 10.1001/archpsyc.1989.01810100009002. [DOI] [PubMed] [Google Scholar]
  • 48.Heston LL. Psychiatric disorders in foster home reared children of schizophrenic mothers. Br J Psychiatry. 1966;112:819–25. doi: 10.1192/bjp.112.489.819. [DOI] [PubMed] [Google Scholar]
  • 49.Kety SS. Schizophrenic illness in the families of schizophrenic adoptees: findings from the Danish national sample. Schizophr Bull. 1988;14:217–22. doi: 10.1093/schbul/14.2.217. [DOI] [PubMed] [Google Scholar]
  • 50.Owen MJ, O’Donovan MC. Schizophrenia genetics. In: Plomin R, Defries JC, Craig IW, et al., editors. Behavioral genetics in the post-genomic era. Washington (DC): American Psychological Association; 2003. pp. 463–82. [Google Scholar]
  • 51.Moises H, Zoega T, Li L, et al. Genes and neurodevelopment in schizophrenia. In: DiLalla LF, editor. Behavior genetics principles: perspectives in development, personality, and psychopathology. Washington (DC): American Psychological Association; 2004. pp. 145–57. [Google Scholar]
  • 52.Neves-Pereira M, Cheung JK, Pasdar A, et al. BDNF gene is a risk factor for schizophrenia in a Scottish population. Mol Psychiatry. 2005;10:208–12. doi: 10.1038/sj.mp.4001575. [DOI] [PubMed] [Google Scholar]
  • 53.Nicholson IR, Neufeld RW. Classification of the schizophrenias according to symptomatology: a two-factor model. J Abnorm Psychol. 1993;102:259–70. doi: 10.1037//0021-843x.102.2.259. [DOI] [PubMed] [Google Scholar]
  • 54.Beauchaine TP. The role of biomarkers and endophenotypes in prevention and treatment of psychopathological disorders. Biomark Med. 2009;3:1–3. doi: 10.2217/17520363.3.1.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Gottesman II, Gould TD. The endophenotype concept in psychiatry: etymology and strategic intentions. Am J Psychiatry. 2003;160:636–45. doi: 10.1176/appi.ajp.160.4.636. [DOI] [PubMed] [Google Scholar]
  • 56.Braff DL, Freedman R, Schork NJ, et al. Deconstructing schizophrenia: an overview of the use of endophenotypes in order to understand a complex disorder. Schizophr Bull. 2007;33:21–32. doi: 10.1093/schbul/sbl049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Cannon TD. The inheritance of intermediate phenotypes for schizophrenia. Curr Opin Psychiatry. 2005;18:135–40. doi: 10.1097/00001504-200503000-00005. [DOI] [PubMed] [Google Scholar]
  • 58.Matthysse S, Holzman PS, Gusella JF, et al. Linkage of eye movement dysfunction to chromosome 6p in schizophrenia: additional evidence. Am J Med Genet B Neuropsychiatr Genet. 2004;128B:30–6. doi: 10.1002/ajmg.b.30030. [DOI] [PubMed] [Google Scholar]
  • 59.Bearden CE, Freimer NB. Endophenotypes for psychiatric disorders: Ready for primetime? Trends Genet. 2006;22:306–13. doi: 10.1016/j.tig.2006.04.004. [DOI] [PubMed] [Google Scholar]
  • 60.Hong LE, Mitchell BD, Avila MT, et al. Familial aggregation of eye-tracking endophenotypes in families of schizophrenic patients. Arch Gen Psychiatry. 2006;63:259–64. doi: 10.1001/archpsyc.63.3.259. [DOI] [PubMed] [Google Scholar]
  • 61.Neufeld RWJ, Williamson P. Neuropsychological correlates of positive symptoms: delusions and hallucinations. In: Pantelis C, Nelson HE, Barnes TRE, editors. Schizophrenia: a neuropsychological perspective. London: John Wiley & Sons; 1996. pp. 205–35. [Google Scholar]
  • 62.Arolt V, Lencer R, Nolte A, et al. Eye tracking dysfunction is a putative phenotypic susceptibility marker of schizophrenia and maps to a locus on chromosome 6p in families with multiple occurrence of the disease. Am J Med Genet. 1996;67:564–79. doi: 10.1002/(SICI)1096-8628(19961122)67:6<564::AID-AJMG10>3.0.CO;2-R. [DOI] [PubMed] [Google Scholar]
  • 63.Lencer R, Malchow CP, Krecker K, et al. Smooth pursuit performance in families with multiple occurrence of schizophrenia and nonpsychotic families. Biol Psychiatry. 1999;45:694–703. doi: 10.1016/s0006-3223(98)00310-2. [DOI] [PubMed] [Google Scholar]
  • 64.Jeon YW, Polich J. Meta-analysis of P300 and schizophrenia: patients, paradigms, and practical implications. Psychophysiology. 2003;40:684–701. doi: 10.1111/1469-8986.00070. [DOI] [PubMed] [Google Scholar]
  • 65.Bramon E, Rabe-Hesketh S, Sham P, et al. Meta-analysis of the P300 and P50 waveforms in schizophrenia. Schizophr Res. 2004;70:315–29. doi: 10.1016/j.schres.2004.01.004. [DOI] [PubMed] [Google Scholar]
  • 66.Bramon E, McDonald C, Croft RJ, et al. Is the P300 wave an endophenotype for schizophrenia? A meta-analysis and a family study. Neuroimage. 2005;27:960–8. doi: 10.1016/j.neuroimage.2005.05.022. [DOI] [PubMed] [Google Scholar]
  • 67.Cadenhead KS, Light GA, Shafer KM, et al. P50 suppression in individuals at risk for schizophrenia: the convergence of clinical, familial, and vulnerability marker risk assessment. Biol Psychiatry. 2005;57:1504–9. doi: 10.1016/j.biopsych.2005.03.003. [DOI] [PubMed] [Google Scholar]
  • 68.Aleman A, Hijman R, de Haan EH, et al. Memory impairment in schizophrenia: a meta-analysis. Am J Psychiatry. 1999;156:1358–66. doi: 10.1176/ajp.156.9.1358. [DOI] [PubMed] [Google Scholar]
  • 69.Toulopoulou T, Rabe-Hesketh S, King H, et al. Episodic memory in schizophrenic patients and their relatives. Schizophr Res. 2003;63:261–71. doi: 10.1016/s0920-9964(02)00324-9. [DOI] [PubMed] [Google Scholar]
  • 70.Braff DL, Geyer MA, Swerdlow NR. Human studies of prepulse inhibition of startle: normal subjects, patient groups, and pharmacological studies. Psychopharmacology (Berl) 2001;156:234–58. doi: 10.1007/s002130100810. [DOI] [PubMed] [Google Scholar]
  • 71.Braff DL, Grillon C, Geyer MA. Gating and habituation of the startle reflex in schizophrenic patients. Arch Gen Psychiatry. 1992;49:206–15. doi: 10.1001/archpsyc.1992.01820030038005. [DOI] [PubMed] [Google Scholar]
  • 72.Broga MI, Neufeld RW. Evaluation of information sequential aspects of schizophrenic performance. I. Framework and current findings. J Nerv Ment Dis. 1981;169:558–68. doi: 10.1097/00005053-198109000-00003. [DOI] [PubMed] [Google Scholar]
  • 73.Roussos P, Giakoumaki SG, Rogdaki M, et al. Prepulse inhibition of the startle reflex depends on the catechol O-methyltransferase Val158Met gene polymorphism. Psychol Med. 2008;38:1651–8. doi: 10.1017/S0033291708002912. [DOI] [PubMed] [Google Scholar]
  • 74.Dawson ME, Schell AM, Hazlett EA, et al. On the clinical and cognitive meaning of impaired sensorimotor gating in schizophrenia. Psychiatry Res. 2000;96:187–97. doi: 10.1016/s0165-1781(00)00208-0. [DOI] [PubMed] [Google Scholar]
  • 75.Braff DL, Swerdlow NR, Geyer MA. Symptom correlates of prepulse inhibition deficits in male schizophrenic patients. Am J Psychiatry. 1999;156:596–602. doi: 10.1176/ajp.156.4.596. [DOI] [PubMed] [Google Scholar]
  • 76.Ludewig K, Geyer MA, Vollenweider FX. Deficits in prepulse inhibition and habituation in never-medicated, first-episode schizophrenia. Biol Psychiatry. 2003;54:121–8. doi: 10.1016/s0006-3223(02)01925-x. [DOI] [PubMed] [Google Scholar]
  • 77.Cadenhead KS, Swerdlow NR, Shafer KM, et al. Modulation of the startle response and startle laterality in relatives of schizophrenic patients and in subjects with schizotypal personality disorder: evidence of inhibitory deficits. Am J Psychiatry. 2000;157:1660–8. doi: 10.1176/appi.ajp.157.10.1660. [DOI] [PubMed] [Google Scholar]
  • 78.Ludewig K, Vollenweider FX. Impaired sensorimotor gating in schizophrenia with deficit and with nondeficit syndrome. Swiss Med Wkly. 2002;132:159–65. doi: 10.4414/smw.2002.09873. [DOI] [PubMed] [Google Scholar]
  • 79.Powell SB, Zhou X, Geyer MA. Prepulse inhibition and genetic mouse models of schizophrenia. Behav Brain Res. 2009;204:282–94. doi: 10.1016/j.bbr.2009.04.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Ellenbroek BA. Pre-attentive processing and schizophrenia: animal studies. Psychopharmacology (Berl) 2004;174:65–74. doi: 10.1007/s00213-003-1684-7. [DOI] [PubMed] [Google Scholar]
  • 81.Baddeley A. Working memory: looking back and looking forward. Nat Rev Neurosci. 2003;4:829–39. doi: 10.1038/nrn1201. [DOI] [PubMed] [Google Scholar]
  • 82.Conklin HM, Curtis CE, Calkins ME, et al. Working memory functioning in schizophrenia patients and their first-degree relatives: cognitive functioning shedding light on etiology. Neuropsychologia. 2005;43:930–42. doi: 10.1016/j.neuropsychologia.2004.09.013. [DOI] [PubMed] [Google Scholar]
  • 83.Pirkola T, Tuulio-Henriksson A, Glahn D, et al. Spatial working memory function in twins with schizophrenia and bipolar disorder. Biol Psychiatry. 2005;58:930–6. doi: 10.1016/j.biopsych.2005.05.041. [DOI] [PubMed] [Google Scholar]
  • 84.Twamley EW, Palmer BW, Jeste DV, et al. Transient and executive function working memory in schizophrenia. Schizophr Res. 2006;87:185–90. doi: 10.1016/j.schres.2006.04.013. [DOI] [PubMed] [Google Scholar]
  • 85.Saperstein AM, Fuller RL, Avila MT, et al. Spatial working memory as a cognitive endophenotype of schizophrenia: assessing risk for pathophysiological dysfunction. Schizophr Bull. 2006;32:498–506. doi: 10.1093/schbul/sbj072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Brébion G, David AS, Jones HM, et al. Working memory span and motor and cognitive speed in schizophrenia. Cogn Behav Neurol. 2009;22:101–8. doi: 10.1097/WNN.0b013e3181a722a0. [DOI] [PubMed] [Google Scholar]
  • 87.O’Connor M, Harris JM, McIntosh AM, et al. Specific cognitive deficits in a group at genetic high risk of schizophrenia. Psychol Med. 2009;39:1649–55. doi: 10.1017/S0033291709005303. [DOI] [PubMed] [Google Scholar]
  • 88.Lee J, Park S. Working memory impairments in schizophrenia: a meta-analysis. J Abnorm Psychol. 2005;114:599–611. doi: 10.1037/0021-843X.114.4.599. [DOI] [PubMed] [Google Scholar]
  • 89.Snitz BE, Macdonald AW, III, Carter CS. Cognitive deficits in unaffected first-degree relatives of schizophrenia patients: a meta-analytic review of putative endophenotypes. Schizophr Bull. 2006;32:179–94. doi: 10.1093/schbul/sbi048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Neufeld RW. On the centrality and significance of stimulus-encoding deficit in schizophrenia. Schizophr Bull. 2007;33:982–93. doi: 10.1093/schbul/sbm056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Mathes B, Wood SJ, Proffitt TM, et al. Early processing deficits in object working memory in first-episode schizophreniform psychosis and established schizophrenia. Psychol Med. 2005;35:1053–62. doi: 10.1017/s0033291705004617. [DOI] [PubMed] [Google Scholar]
  • 92.Neufeld RWJ, Boksman K, Vollick D, et al. Stochastic dynamics of stimulus encoding in schizophrenia: theory, testing, and application. J Math Psychol. 2010;54:90–108. [Google Scholar]
  • 93.Rodriguez-Sanchez JM, Crespo-Facorro B, Gonzalez-Blanch C, et al. Cognitive dysfunction in first-episode psychosis: the processing speed hypothesis. Br J Psychiatry Suppl. 2007;51:s107–s110. doi: 10.1192/bjp.191.51.s107. [DOI] [PubMed] [Google Scholar]
  • 94.Morris R. Developments of a water-maze procedure for studying spatial learning in the rat. J Neurosci Methods. 1984;11:47–60. doi: 10.1016/0165-0270(84)90007-4. [DOI] [PubMed] [Google Scholar]
  • 95.Deacon RM, Rawlins JN. Learning impairments of hippocampal-lesioned mice in a paddling pool. Behav Neurosci. 2002;116:472–8. doi: 10.1037//0735-7044.116.3.472. [DOI] [PubMed] [Google Scholar]
  • 96.Morris RG, Hagan JJ, Rawlins JN. Allocentric spatial learning by hippocampectomised rats: a further test of the “spatial mapping” and “working memory” theories of hippocampal function. Q J Exp Psychol B. 1986;38:365–95. [PubMed] [Google Scholar]
  • 97.Sandstrom NJ, Kim JH, Wasserman MA. Testosterone modulates performance on a spatial working memory task in male rats. Horm Behav. 2006;50:18–26. doi: 10.1016/j.yhbeh.2005.09.008. [DOI] [PubMed] [Google Scholar]
  • 98.Koike H, Arguello PA, Kvajo M, et al. Disc1 is mutated in the 129S6/SvEv strain and modulates working memory in mice. Proc Natl Acad Sci U S A. 2006;103:3693–7. doi: 10.1073/pnas.0511189103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Castner SA, Goldman-Rakic PS, Williams GV. Animal models of working memory: insights for targeting cognitive dysfunction in schizophrenia. Psychopharmacology (Berl) 2004;174:111–25. doi: 10.1007/s00213-003-1710-9. [DOI] [PubMed] [Google Scholar]
  • 100.Ceaser AE, Goldberg TE, Egan MF, et al. Set-shifting ability and schizophrenia: A marker of clinical illness or an intermediate phenotype? Biol Psychiatry. 2008;64:782–8. doi: 10.1016/j.biopsych.2008.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.McKirdy J, Sussmann JE, Hall J, et al. Set shifting and reversal learning in patients with bipolar disorder or schizophrenia. Psychol Med. 2009;39:1289–93. doi: 10.1017/S0033291708004935. [DOI] [PubMed] [Google Scholar]
  • 102.Morrens M, Hulstijn W, Lewi PJ, et al. Stereotypy in schizophrenia. Schizophr Res. 2006;84:397–404. doi: 10.1016/j.schres.2006.01.024. [DOI] [PubMed] [Google Scholar]
  • 103.Yogev H, Hadar U, Gutman Y, et al. Perseveration and over-switching in schizophrenia. Schizophr Res. 2003;61:315–21. doi: 10.1016/s0920-9964(02)00323-7. [DOI] [PubMed] [Google Scholar]
  • 104.Franke P, Maier W, Hardt J, et al. Cognitive functioning and anhedonia in subjects at risk for schizophrenia. Schizophr Res. 1993;10:77–84. doi: 10.1016/0920-9964(93)90079-x. [DOI] [PubMed] [Google Scholar]
  • 105.Szöke A, Schurhoff F, Mathieu F, et al. Tests of executive functions in first-degree relatives of schizophrenic patients: a meta-analysis. Psychol Med. 2005;35:771–82. doi: 10.1017/s0033291704003460. [DOI] [PubMed] [Google Scholar]
  • 106.Geyer MA, Moghaddam B. Animal models relevant to schizophrenia disorders. In: Davis KL, Charney D, Coyle JT, et al., editors. Neuropsychopharmacology: the fifth generation of progress. Philadelphia (PA): Lippincott, Williams and Wilkins; 2002. pp. 690–701. [Google Scholar]
  • 107.Sams-Dodd F. Distinct effects of d-amphetamine and phencyclidine on the social behaviour of rats. Behav Pharmacol. 1995;6:55–65. [PubMed] [Google Scholar]
  • 108.Sams-Dodd F. Effects of continuous D-amphetamine and phencyclidine administration on social behaviour, stereotyped behaviour, and locomotor activity in rats. Neuropsychopharmacology. 1998;19:18–25. doi: 10.1016/S0893-133X(97)00200-5. [DOI] [PubMed] [Google Scholar]
  • 109.Weiner I. The “two-headed” latent inhibition model of schizophrenia: modeling positive and negative symptoms and their treatment. Psychopharmacology (Berl) 2003;169:257–97. doi: 10.1007/s00213-002-1313-x. [DOI] [PubMed] [Google Scholar]
  • 110.Vaitl D, Lipp O, Bauer U, et al. Latent inhibition and schizophrenia: Pavlovian conditioning of autonomic responses. Schizophr Res. 2002;55:147–58. doi: 10.1016/s0920-9964(01)00250-x. [DOI] [PubMed] [Google Scholar]
  • 111.Gal G, Barnea Y, Biran L, et al. Enhancement of latent inhibition in patients with chronic schizophrenia. Behav Brain Res. 2009;197:1–8. doi: 10.1016/j.bbr.2008.08.023. [DOI] [PubMed] [Google Scholar]
  • 112.Gray NS, Pilowsky LS, Gray JA, et al. Latent inhibition in drug naive schizophrenics: relationship to duration of illness and dopamine D2 binding using SPET. Schizophr Res. 1995;17:95–107. doi: 10.1016/0920-9964(95)00034-j. [DOI] [PubMed] [Google Scholar]
  • 113.Martins Serra A, Jones SH, Toone B, et al. Impaired associative learning in chronic schizophrenics and their first-degree relatives: a study of latent inhibition and the Kamin blocking effect. Schizophr Res. 2001;48:273–89. doi: 10.1016/s0920-9964(00)00141-9. [DOI] [PubMed] [Google Scholar]
  • 114.Gray NS, Snowden RJ. The relevance of irrelevance to schizophrenia. Neurosci Biobehav Rev. 2005;29:989–99. doi: 10.1016/j.neubiorev.2005.01.006. [DOI] [PubMed] [Google Scholar]
  • 115.Swerdlow NR, Stephany N, Wasserman LC, et al. Intact visual latent inhibition in schizophrenia patients in a within-subject paradigm. Schizophr Res. 2005;72:169–83. doi: 10.1016/j.schres.2004.03.022. [DOI] [PubMed] [Google Scholar]
  • 116.Bediou B, Franck N, Saoud M, et al. Effects of emotion and identity on facial affect processing in schizophrenia. Psychiatry Res. 2005;133:149–57. doi: 10.1016/j.psychres.2004.08.008. [DOI] [PubMed] [Google Scholar]
  • 117.Bellack AS, Morrison RL, Wixted JT, et al. An analysis of social competence in schizophrenia. Br J Psychiatry. 1990;156:809–18. doi: 10.1192/bjp.156.6.809. [DOI] [PubMed] [Google Scholar]
  • 118.Halford WK, Hayes RL. Social skills in schizophrenia: assessing the relationship between social skills, psychopathology and community functioning. Soc Psychiatry Psychiatr Epidemiol. 1995;30:14–9. doi: 10.1007/BF00784429. [DOI] [PubMed] [Google Scholar]
  • 119.Penn DL, Corrigan PW, Bentall RP, et al. Social cognition in schizophrenia. Psychol Bull. 1997;121:114–32. doi: 10.1037/0033-2909.121.1.114. [DOI] [PubMed] [Google Scholar]
  • 120.Craver JC, Pogue-Geile MF. Familial liability to schizophrenia: a sibling study of negative symptoms. Schizophr Bull. 1999;25:827–39. doi: 10.1093/oxfordjournals.schbul.a033422. [DOI] [PubMed] [Google Scholar]
  • 121.Katsanis J, Iacono WG, Beiser M. Anhedonia and perceptual aberration in first-episode psychotic patients and their relatives. J Abnorm Psychol. 1990;99:202–6. doi: 10.1037//0021-843x.99.2.202. [DOI] [PubMed] [Google Scholar]
  • 122.Miyakawa T, Leiter LM, Gerber DJ, et al. Conditional calcineurin knockout mice exhibit multiple abnormal behaviors related to schizophrenia. Proc Natl Acad Sci U S A. 2003;100:8987–92. doi: 10.1073/pnas.1432926100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Crawley JN. Designing mouse behavioral tasks relevant to autistic-like behaviors. Ment Retard Dev Disabil Res Rev. 2004;10:248–58. doi: 10.1002/mrdd.20039. [DOI] [PubMed] [Google Scholar]
  • 124.Sams-Dodd F. Phencyclidine-induced stereotyped behaviour and social isolation in rats: a possible animal model of schizophrenia. Behav Pharmacol. 1996;7:3–23. [PubMed] [Google Scholar]
  • 125.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;132:303–10. doi: 10.1007/s002130050349. [DOI] [PubMed] [Google Scholar]
  • 126.Powell CM, Miyakawa T. Schizophrenia-relevant behavioral testing in rodent models: A uniquely human disorder? Biol Psychiatry. 2006;59:1198–207. doi: 10.1016/j.biopsych.2006.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Kas MJ, Gelegen C, Schalkwyk LC, et al. Interspecies comparisons of functional genetic variations and their implications in neuropsychiatry. Am J Med Genet B Neuropsychiatr Genet. 2009;150B:309–17. doi: 10.1002/ajmg.b.30815. [DOI] [PubMed] [Google Scholar]
  • 128.Ellenbroek BA, Cools AR. Animal models for the negative symptoms of schizophrenia. Behav Pharmacol. 2000;11:223–33. doi: 10.1097/00008877-200006000-00006. [DOI] [PubMed] [Google Scholar]
  • 129.McDannald M, Schoenbaum G. Toward a model of impaired reality testing in rats. Schizophr Bull. 2009;35:664–7. doi: 10.1093/schbul/sbp050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Marcotte ER, Pearson DM, Srivastava LK. Animal models of schizophrenia: a critical review. J Psychiatry Neurosci. 2001;26:395–410. [PMC free article] [PubMed] [Google Scholar]
  • 131.Shultz SR, Macfabe DF, Ossenkopp KP, et al. Intracerebroventricular injection of propionic acid, an enteric bacterial metabolic end-product, impairs social behavior in the rat: implications for an animal model of autism. Neuropharmacology. 2008;54:901–11. doi: 10.1016/j.neuropharm.2008.01.013. [DOI] [PubMed] [Google Scholar]
  • 132.Rogers DC, Jones DN, Nelson PR, et al. Use of SHIRPA and discriminant analysis to characterise marked differences in the behavioural phenotype of six inbred mouse strains. Behav Brain Res. 1999;105:207–17. doi: 10.1016/s0166-4328(99)00072-8. [DOI] [PubMed] [Google Scholar]
  • 133.Ellenbroek BA. Animal models in the genomic era: possibilities and limitations with special emphasis on schizophrenia. Behav Pharmacol. 2003;14:409–17. doi: 10.1097/01.fbp.0000087736.21047.3c. [DOI] [PubMed] [Google Scholar]
  • 134.Ellison GD, Eison MS. Continuous amphetamine intoxication: an animal model of the acute psychotic episode. Psychol Med. 1983;13:751–61. doi: 10.1017/s003329170005145x. [DOI] [PubMed] [Google Scholar]
  • 135.Mandillo S, Rinaldi A, Oliverio A, et al. Repeated administration of phencyclidine, amphetamine and MK-801 selectively impairs spatial learning in mice: a possible model of psychotomimetic drug-induced cognitive deficits. Behav Pharmacol. 2003;14:533–44. doi: 10.1097/00008877-200311000-00006. [DOI] [PubMed] [Google Scholar]
  • 136.Mansbach RS, Geyer MA, Braff DL. Dopaminergic stimulation disrupts sensorimotor gating in the rat. Psychopharmacology (Berl) 1988;94:507–14. doi: 10.1007/BF00212846. [DOI] [PubMed] [Google Scholar]
  • 137.Steinpreis RE, Sokolowski JD, Papanikolaou A, et al. The effects of haloperidol and clozapine on PCP- and amphetamine-induced suppression of social behavior in the rat. Pharmacol Biochem Behav. 1994;47:579–85. doi: 10.1016/0091-3057(94)90162-7. [DOI] [PubMed] [Google Scholar]
  • 138.Ellenbroek BA, Budde S, Cools AR. Prepulse inhibition and latent inhibition: the role of dopamine in the medial prefrontal cortex. Neuroscience. 1996;75:535–42. doi: 10.1016/0306-4522(96)00307-7. [DOI] [PubMed] [Google Scholar]
  • 139.Andersen JD, Pouzet B. Spatial memory deficits induced by perinatal treatment of rats with PCP and reversal effect of D-serine. Neuropsychopharmacology. 2004;29:1080–90. doi: 10.1038/sj.npp.1300394. [DOI] [PubMed] [Google Scholar]
  • 140.Rasmussen BA, O’Neil J, Manaye KF, et al. Long-term effects of developmental PCP administration on sensorimotor gating in male and female rats. Psychopharmacology (Berl) 2007;190:43–9. doi: 10.1007/s00213-006-0584-z. [DOI] [PubMed] [Google Scholar]
  • 141.Sams-Dodd F. Phencyclidine in the social interaction test: an animal model of schizophrenia with face and predictive validity. Rev Neurosci. 1999;10:59–90. doi: 10.1515/revneuro.1999.10.1.59. [DOI] [PubMed] [Google Scholar]
  • 142.Brigman JL, Ihne J, Saksida LM, et al. Effects of subchronic phencyclidine (PCP) treatment on social behaviors, and operant discrimination and reversal learning in C57BL/6J mice. Front Behav Neurosci. 2009;3:2. doi: 10.3389/neuro.08.002.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Liu J, Suzuki T, Seki T, et al. Effects of repeated phencyclidine administration on adult hippocampal neurogenesis in the rat. Synapse. 2006;60:56–68. doi: 10.1002/syn.20275. [DOI] [PubMed] [Google Scholar]
  • 144.Sams-Dodd F. Effect of novel antipsychotic drugs on phencyclidine-induced stereotyped behaviour and social isolation in the rat social interaction test. Behav Pharmacol. 1997;8:196–215. [PubMed] [Google Scholar]
  • 145.Sams-Dodd F. Effects of diazepam, citalopram, methadone and naloxone on PCP-induced stereotyped behaviour and social isolation in the rat social interaction test. Neurosci Biobehav Rev. 1998;23:287–93. doi: 10.1016/s0149-7634(98)00030-x. [DOI] [PubMed] [Google Scholar]
  • 146.Becker A, Peters B, Schroeder H, et al. Ketamine-induced changes in rat behaviour: a possible animal model of schizophrenia. Prog Neuropsychopharmacol Biol Psychiatry. 2003;27:687–700. doi: 10.1016/S0278-5846(03)00080-0. [DOI] [PubMed] [Google Scholar]
  • 147.Becker A, Grecksch G. Ketamine-induced changes in rat behaviour: a possible animal model of schizophrenia. Test of predictive validity. Prog Neuropsychopharmacol Biol Psychiatry. 2004;28:1267–77. doi: 10.1016/j.pnpbp.2004.06.019. [DOI] [PubMed] [Google Scholar]
  • 148.Baier PC, Blume A, Koch J, et al. Early postnatal depletion of NMDA receptor development affects behaviour and NMDA receptor expression until later adulthood in rats — a possible model for schizophrenia. Behav Brain Res. 2009;205:96–101. doi: 10.1016/j.bbr.2009.06.018. [DOI] [PubMed] [Google Scholar]
  • 149.Japha K, Koch M. Picrotoxin in the medial prefrontal cortex impairs sensorimotor gating in rats: reversal by haloperidol. Psychopharmacology (Berl) 1999;144:347–54. doi: 10.1007/s002130051017. [DOI] [PubMed] [Google Scholar]
  • 150.Berretta S, Munno DW, Benes FM. Amygdalar activation alters the hippocampal GABA system: “partial” modelling for postmortem changes in schizophrenia. J Comp Neurol. 2001;431:129–38. doi: 10.1002/1096-9861(20010305)431:2<129::aid-cne1060>3.0.co;2-6. [DOI] [PubMed] [Google Scholar]
  • 151.Berretta S, Lange N, Bhattacharyya S, et al. Long-term effects of amygdala GABA receptor blockade on specific subpopulations of hippocampal interneurons. Hippocampus. 2004;14:876–94. doi: 10.1002/hipo.20002. [DOI] [PubMed] [Google Scholar]
  • 152.Gisabella B, Cunningham MG, Bolshakov VY, et al. Amygdala-dependent regulation of electrical properties of hippocampal interneurons in a model of schizophrenia. Biol Psychiatry. 2009;65:464–72. doi: 10.1016/j.biopsych.2008.09.016. [DOI] [PubMed] [Google Scholar]
  • 153.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]
  • 154.Paz RD, Tardito S, Atzori M, et al. Glutamatergic dysfunction in schizophrenia: from basic neuroscience to clinical psychopharmacology. Eur Neuropsychopharmacol. 2008;18:773–86. doi: 10.1016/j.euroneuro.2008.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Gainetdinov RR, Mohn AR, Caron MG. Genetic animal models: focus on schizophrenia. Trends Neurosci. 2001;24:527–33. doi: 10.1016/s0166-2236(00)01886-5. [DOI] [PubMed] [Google Scholar]
  • 156.Ralph RJ, Paulus MP, Fumagalli F, et al. Prepulse inhibition deficits and perseverative motor patterns in dopamine transporter knock-out mice: differential effects of D1 and D2 receptor antagonists. J Neurosci. 2001;21:305–13. doi: 10.1523/JNEUROSCI.21-01-00305.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Spielewoy C, Roubert C, Hamon M, et al. Behavioural disturbances associated with hyperdopaminergia in dopamine-transporter knockout mice. Behav Pharmacol. 2000;11:279–90. doi: 10.1097/00008877-200006000-00011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Fradley RL, O’Meara GF, Newman RJ, et al. STOP knockout and NMDA NR1 hypomorphic mice exhibit deficits in sensorimotor gating. Behav Brain Res. 2005;163:257–64. doi: 10.1016/j.bbr.2005.05.012. [DOI] [PubMed] [Google Scholar]
  • 159.Halene TB, Ehrlichman RS, Liang Y, et al. Assessment of NMDA receptor NR1 subunit hypofunction in mice as a model for schizophrenia. Genes Brain Behav. 2009;8:661–75. doi: 10.1111/j.1601-183X.2009.00504.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Mohn AR, Gainetdinov RR, Caron MG, et al. Mice with reduced NMDA receptor expression display behaviors related to schizophrenia. Cell. 1999;98:427–36. doi: 10.1016/s0092-8674(00)81972-8. [DOI] [PubMed] [Google Scholar]
  • 161.Chen J, Lipska BK, Weinberger DR. Genetic mouse models of schizophrenia: from hypothesis-based to susceptibility gene-based models. Biol Psychiatry. 2006;59:1180–8. doi: 10.1016/j.biopsych.2006.02.024. [DOI] [PubMed] [Google Scholar]
  • 162.Brown AS, Begg MD, Gravenstein S, et al. Serologic evidence of prenatal influenza in the etiology of schizophrenia. Arch Gen Psychiatry. 2004;61:774–80. doi: 10.1001/archpsyc.61.8.774. [DOI] [PubMed] [Google Scholar]
  • 163.Kendell RE, Kemp IW. Maternal influenza in the etiology of schizophrenia. Arch Gen Psychiatry. 1989;46:878–82. doi: 10.1001/archpsyc.1989.01810100020004. [DOI] [PubMed] [Google Scholar]
  • 164.Brown AS. Prenatal infection as a risk factor for schizophrenia. Schizophr Bull. 2006;32:200–2. doi: 10.1093/schbul/sbj052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Fatemi SH, Reutiman TJ, Folsom TD, et al. Maternal infection leads to abnormal gene regulation and brain atrophy in mouse offspring: implications for genesis of neurodevelopmental disorders. Schizophr Res. 2008;99:56–70. doi: 10.1016/j.schres.2007.11.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Fatemi SH, Pearce DA, Brooks AI, et al. Prenatal viral infection in mouse causes differential expression of genes in brains of mouse progeny: a potential animal model for schizophrenia and autism. Synapse. 2005;57:91–9. doi: 10.1002/syn.20162. [DOI] [PubMed] [Google Scholar]
  • 167.Shi L, Fatemi SH, Sidwell RW, et al. Maternal influenza infection causes marked behavioral and pharmacological changes in the offspring. J Neurosci. 2003;23:297–302. doi: 10.1523/JNEUROSCI.23-01-00297.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Boksa P. Maternal infection during pregnancy and schizophrenia. J Psychiatry Neurosci. 2008;33:183–5. [PMC free article] [PubMed] [Google Scholar]
  • 169.Boksa P. Animal models of obstetric complications in relation to schizophrenia. Brain Res Brain Res Rev. 2004;45:1–17. doi: 10.1016/j.brainresrev.2004.01.001. [DOI] [PubMed] [Google Scholar]
  • 170.Engel JA, Zhang J, Bergstrom T, et al. Neonatal herpes simplex virus type 1 brain infection affects the development of sensorimotor gating in rats. Brain Res. 2000;863:233–40. doi: 10.1016/s0006-8993(00)02149-1. [DOI] [PubMed] [Google Scholar]
  • 171.Pletnikov MV, Rubin SA, Vogel MW, et al. Effects of genetic background on neonatal Borna disease virus infection-induced neurodevelopmental damage. II. Neurochemical alterations and responses to pharmacological treatments. Brain Res. 2002;944:108–23. doi: 10.1016/s0006-8993(02)02724-5. [DOI] [PubMed] [Google Scholar]
  • 172.Ashdown H, Dumont Y, Ng M, et al. The role of cytokines in mediating effects of prenatal infection on the fetus: implications for schizophrenia. Mol Psychiatry. 2006;11:47–55. doi: 10.1038/sj.mp.4001748. [DOI] [PubMed] [Google Scholar]
  • 173.Shi L, Tu N, Patterson PH. Maternal influenza infection is likely to alter fetal brain development indirectly: the virus is not detected in the fetus. Int J Dev Neurosci. 2005;23:299–305. doi: 10.1016/j.ijdevneu.2004.05.005. [DOI] [PubMed] [Google Scholar]
  • 174.Meyer U, Feldon J, Schedlowski M, et al. Immunological stress at the maternal-foetal interface: a link between neurodevelopment and adult psychopathology. Brain Behav Immun. 2006;20:378–88. doi: 10.1016/j.bbi.2005.11.003. [DOI] [PubMed] [Google Scholar]
  • 175.Nawa H, Takei N. Recent progress in animal modeling of immune inflammatory processes in schizophrenia: implication of specific cytokines. Neurosci Res. 2006;56:2–13. doi: 10.1016/j.neures.2006.06.002. [DOI] [PubMed] [Google Scholar]
  • 176.Patterson PH. Immune involvement in schizophrenia and autism: etiology, pathology and animal models. Behav Brain Res. 2009;204:313–21. doi: 10.1016/j.bbr.2008.12.016. [DOI] [PubMed] [Google Scholar]
  • 177.Dalman C, Allebeck P, Cullberg J, et al. Obstetric complications and the risk of schizophrenia: a longitudinal study of a national birth cohort. Arch Gen Psychiatry. 1999;56:234–40. doi: 10.1001/archpsyc.56.3.234. [DOI] [PubMed] [Google Scholar]
  • 178.Geddes JR, Lawrie SM. Obstetric complications and schizophrenia: a meta-analysis. Br J Psychiatry. 1995;167:786–93. doi: 10.1192/bjp.167.6.786. [DOI] [PubMed] [Google Scholar]
  • 179.Avila MT, Sherr J, Valentine LE, et al. Neurodevelopmental interactions conferring risk for schizophrenia: a study of dermatoglyphic markers in patients and relatives. Schizophr Bull. 2003;29:595–605. doi: 10.1093/oxfordjournals.schbul.a007031. [DOI] [PubMed] [Google Scholar]
  • 180.Rapoport JL, Addington AM, Frangou S, et al. The neurodevelopmental model of schizophrenia: update 2005. Mol Psychiatry. 2005;10:434–49. doi: 10.1038/sj.mp.4001642. [DOI] [PubMed] [Google Scholar]
  • 181.Nicodemus KK, Marenco S, Batten AJ, et al. Serious obstetric complications interact with hypoxia-regulated/vascular-expression genes to influence schizophrenia risk. Mol Psychiatry. 2008;13:873–7. doi: 10.1038/sj.mp.4002153. [DOI] [PubMed] [Google Scholar]
  • 182.Nicholson IR, Neufeld RW. A dynamic vulnerability perspective on stress and schizophrenia. Am J Orthopsychiatry. 1992;62:117–30. doi: 10.1037/h0079307. [DOI] [PubMed] [Google Scholar]
  • 183.Boksa P, El-Khodor BF. Birth insult interacts with stress at adulthood to alter dopaminergic function in animal models: possible implications for schizophrenia and other disorders. Neurosci Biobehav Rev. 2003;27:91–101. doi: 10.1016/s0149-7634(03)00012-5. [DOI] [PubMed] [Google Scholar]
  • 184.Laviola G, Adriani W, Rea M, et al. Social withdrawal, neophobia, and stereotyped behavior in developing rats exposed to neonatal asphyxia. Psychopharmacology (Berl) 2004;175:196–205. doi: 10.1007/s00213-004-1800-3. [DOI] [PubMed] [Google Scholar]
  • 185.Tan YL, Zhou DF, Cao LY, et al. Decreased BDNF in serum of patients with chronic schizophrenia on long-term treatment with antipsychotics. Neurosci Lett. 2005;382:27–32. doi: 10.1016/j.neulet.2005.02.054. [DOI] [PubMed] [Google Scholar]
  • 186.Toyooka K, Asama K, Watanabe Y, et al. Decreased levels of brain-derived neurotrophic factor in serum of chronic schizophrenic patients. Psychiatry Res. 2002;110:249–57. doi: 10.1016/s0165-1781(02)00127-0. [DOI] [PubMed] [Google Scholar]
  • 187.Kohlhauser C, Mosgoeller W, Hoeger H, et al. Cholinergic, monoaminergic and glutamatergic changes following perinatal asphyxia in the rat. Cell Mol Life Sci. 1999;55:1491–501. doi: 10.1007/s000180050388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Berger MA, Barros VG, Sarchi MI, et al. Long-term effects of prenatal stress on dopamine and glutamate receptors in adult rat brain. Neurochem Res. 2002;27:1525–33. doi: 10.1023/a:1021656607278. [DOI] [PubMed] [Google Scholar]
  • 189.Kofman O. The role of prenatal stress in the etiology of developmental behavioural disorders. Neurosci Biobehav Rev. 2002;26:457–70. doi: 10.1016/s0149-7634(02)00015-5. [DOI] [PubMed] [Google Scholar]
  • 190.Ellenbroek BA, Derks N, Park HJ. Early maternal deprivation retards neurodevelopment in Wistar rats. Stress. 2005;8:247–57. doi: 10.1080/10253890500404634. [DOI] [PubMed] [Google Scholar]
  • 191.Garner B, Wood SJ, Pantelis C, et al. Early maternal deprivation reduces prepulse inhibition and impairs spatial learning ability in adulthood: no further effect of post-pubertal chronic corticosterone treatment. Behav Brain Res. 2007;176:323–32. doi: 10.1016/j.bbr.2006.10.020. [DOI] [PubMed] [Google Scholar]
  • 192.Lipska BK. Using animal models to test a neurodevelopmental hypothesis of schizophrenia. J Psychiatry Neurosci. 2004;29:282–6. [PMC free article] [PubMed] [Google Scholar]
  • 193.Flores G, Silva-Gomez AB, Ibanez O, et al. Comparative behavioral changes in postpubertal rats after neonatal excitotoxic lesions of the ventral hippocampus and the prefrontal cortex. Synapse. 2005;56:147–53. doi: 10.1002/syn.20140. [DOI] [PubMed] [Google Scholar]
  • 194.Becker A, Grecksch G, Bernstein HG, et al. Social behaviour in rats lesioned with ibotenic acid in the hippocampus: quantitative and qualitative analysis. Psychopharmacology (Berl) 1999;144:333–8. doi: 10.1007/s002130051015. [DOI] [PubMed] [Google Scholar]
  • 195.Grecksch G, Bernstein HG, Becker A, et al. Disruption of latent inhibition in rats with postnatal hippocampal lesions. Neuropsychopharmacology. 1999;20:525–32. doi: 10.1016/S0893-133X(98)00081-5. [DOI] [PubMed] [Google Scholar]
  • 196.Lipska BK, Swerdlow NR, Geyer MA, et al. Neonatal excitotoxic hippocampal damage in rats causes post-pubertal changes in prepulse inhibition of startle and its disruption by apomorphine. Psychopharmacology (Berl) 1995;122:35–43. doi: 10.1007/BF02246439. [DOI] [PubMed] [Google Scholar]
  • 197.Lipska BK, Aultman JM, Verma A, et al. Neonatal damage of the ventral hippocampus impairs working memory in the rat. Neuropsychopharmacology. 2002;27:47–54. doi: 10.1016/S0893-133X(02)00282-8. [DOI] [PubMed] [Google Scholar]
  • 198.Schroeder H, Grecksch G, Becker A, et al. Alterations of the dopaminergic and glutamatergic neurotransmission in adult rats with postnatal ibotenic acid hippocampal lesion. Psychopharmacology (Berl) 1999;145:61–6. doi: 10.1007/s002130051032. [DOI] [PubMed] [Google Scholar]
  • 199.Tseng KY, Lewis BL, Hashimoto T, et al. A neonatal ventral hippocampal lesion causes functional deficits in adult prefrontal cortical interneurons. J Neurosci. 2008;28:12691–9. doi: 10.1523/JNEUROSCI.4166-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.François J, Ferrandon A, Koning E, et al. Selective reorganization of GABAergic transmission in neonatal ventral hippocampal-lesioned rats. Int J Neuropsychopharmacol. 2009;12:1097–110. doi: 10.1017/S1461145709009985. [DOI] [PubMed] [Google Scholar]
  • 201.Tseng KY, Chambers RA, Lipska BK. The neonatal ventral hippocampal lesion as a heuristic neurodevelopmental model of schizophrenia. Behav Brain Res. 2009;204:295–305. doi: 10.1016/j.bbr.2008.11.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Schneider M, Koch M. Behavioral and morphological alterations following neonatal excitotoxic lesions of the medial prefrontal cortex in rats. Exp Neurol. 2005;195:185–98. doi: 10.1016/j.expneurol.2005.04.014. [DOI] [PubMed] [Google Scholar]
  • 203.Schneider M, Koch M. Deficient social and play behavior in juvenile and adult rats after neonatal cortical lesion: effects of chronic pubertal cannabinoid treatment. Neuropsychopharmacology. 2005;30:944–57. doi: 10.1038/sj.npp.1300634. [DOI] [PubMed] [Google Scholar]
  • 204.Schneider M, Koch M. The effect of chronic peripubertal cannabinoid treatment on deficient object recognition memory in rats after neonatal mPFC lesion. Eur Neuropsychopharmacol. 2007;17:180–6. doi: 10.1016/j.euroneuro.2006.03.009. [DOI] [PubMed] [Google Scholar]
  • 205.Bennay M, Gernert M, Schwabe K, et al. Neonatal medial prefrontal cortex lesion enhances the sensitivity of the mesoaccumbal dopamine system. Eur J Neurosci. 2004;19:3277–90. doi: 10.1111/j.0953-816X.2004.03442.x. [DOI] [PubMed] [Google Scholar]
  • 206.Talamini LM, Koch T, Ter Horst GJ, et al. Methylazoxymethanol acetate-induced abnormalities in the entorhinal cortex of the rat; parallels with morphological findings in schizophrenia. Brain Res. 1998;789:293–306. doi: 10.1016/s0006-8993(98)00020-1. [DOI] [PubMed] [Google Scholar]
  • 207.Fiore M, Korf J, Angelucci F, et al. Prenatal exposure to methylazoxymethanol acetate in the rat alters neurotrophin levels and behavior: considerations for neurodevelopmental diseases. Physiol Behav. 2000;71:57–67. doi: 10.1016/s0031-9384(00)00310-3. [DOI] [PubMed] [Google Scholar]
  • 208.Talamini LM, Koch T, Luiten PG, et al. Interruptions of early cortical development affect limbic association areas and social behaviour in rats; possible relevance for neurodevelopmental disorders. Brain Res. 1999;847:105–20. doi: 10.1016/s0006-8993(99)02067-3. [DOI] [PubMed] [Google Scholar]
  • 209.Flagstad P, Glenthoj BY, Didriksen M. Cognitive deficits caused by late gestational disruption of neurogenesis in rats: a preclinical model of schizophrenia. Neuropsychopharmacology. 2005;30:250–60. doi: 10.1038/sj.npp.1300625. [DOI] [PubMed] [Google Scholar]
  • 210.Flagstad P, Mork A, Glenthoj BY, et al. 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;29:2052–64. doi: 10.1038/sj.npp.1300516. [DOI] [PubMed] [Google Scholar]
  • 211.Moore H, Jentsch JD, Ghajarnia M, et al. A neurobehavioral systems analysis of adult rats exposed to methylazoxymethanol acetate on E17: implications for the neuropathology of schizophrenia. Biol Psychiatry. 2006;60:253–64. doi: 10.1016/j.biopsych.2006.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Futamura T, Kakita A, Tohmi M, et al. Neonatal perturbation of neurotrophic signaling results in abnormal sensorimotor gating and social interaction in adults: implication for epidermal growth factor in cognitive development. Mol Psychiatry. 2003;8:19–29. doi: 10.1038/sj.mp.4001138. [DOI] [PubMed] [Google Scholar]
  • 213.Sotoyama H, Namba H, Takei N, et al. Neonatal exposure to epidermal growth factor induces dopamine D2-like receptor supersensitivity in adult sensorimotor gating. Psychopharmacology (Berl) 2007;191:783–92. doi: 10.1007/s00213-006-0595-9. [DOI] [PubMed] [Google Scholar]
  • 214.Rajakumar N, Leung LS, Ma J, et al. Altered neurotrophin receptor function in the developing prefrontal cortex leads to adult-onset dopaminergic hyperresponsivity and impaired prepulse inhibition of acoustic startle. Biol Psychiatry. 2004;55:797–803. doi: 10.1016/j.biopsych.2003.12.015. [DOI] [PubMed] [Google Scholar]
  • 215.Rajakumar N, Rajakumar B. A mild disruption of subplate function in the developing prefrontal cortex is sufficient to cause multiple neuropathological features of schizophrenia. Biennial Meeting of the International Society for Schizophrenia Research; April 2005; Savannah, Ga.. [Google Scholar]
  • 216.Lazar NL, Rajakumar N, Cain DP. Injections of NGF into neonatal frontal cortex decrease social interaction as adults: a rat model of schizophrenia. Schizophr Bull. 2008;34:127–36. doi: 10.1093/schbul/sbm039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Lavin A, Moore HM, Grace AA. Prenatal disruption of neocortical development alters prefrontal cortical neuron responses to dopamine in adult rats. Neuropsychopharmacology. 2005;30:1426–35. doi: 10.1038/sj.npp.1300696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Sham P, McGiffin P. Linkage and association. In: McGuffin P, Owen MJ, Gottesman II, editors. Psychiatric genetics and genomics. New York (NY): Oxford University Press; 2001. pp. 55–76. [Google Scholar]
  • 219.Gottesman II, Moldin SO. Genotyoes, genes, genesis, and pathogenesis in schizophrenia. In: Lenzenweger MF, Dworkin RH, editors. Origins and development of schizophrenia: advances in experimental psychopathology. Washington (DC): American Psychological Association; 1998. pp. 5–26. [Google Scholar]
  • 220.Badner JA, Gershon ES. Meta-analysis of whole-genome linkage scans of bipolar disorder and schizophrenia. Mol Psychiatry. 2002;7:405–11. doi: 10.1038/sj.mp.4001012. [DOI] [PubMed] [Google Scholar]
  • 221.Lewis CM, Levinson DF, Wise LH, et al. Genome scan meta-analysis of schizophrenia and bipolar disorder, part II: schizophrenia. Am J Hum Genet. 2003;73:34–48. doi: 10.1086/376549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Holmans PA, Riley B, Pulver AE, et al. Genomewide linkage scan of schizophrenia in a large multicenter pedigree sample using single nucleotide polymorphisms. Mol Psychiatry. 2009;14:786–95. doi: 10.1038/mp.2009.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Sobin C, Kiley-Brabeck K, Karayiorgou M. Lower prepulse inhibition in children with the 22q11 deletion syndrome. Am J Psychiatry. 2005;162:1090–9. doi: 10.1176/appi.ajp.162.6.1090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Sobin C, Kiley-Brabeck K, Daniels S, et al. Neuropsychological characteristics of children with the 22q11 deletion syndrome: a descriptive analysis. Child Neuropsychol. 2005;11:39–53. doi: 10.1080/09297040590911167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Jolin EM, Weller RA, Weller EB. Psychosis in children with velocardiofacial syndrome (22q11.2 deletion syndrome) Curr Psychiatry Rep. 2009;11:99–105. doi: 10.1007/s11920-009-0016-y. [DOI] [PubMed] [Google Scholar]
  • 226.Karayiorgou M, Gogos JA. The molecular genetics of the 22q11-associated schizophrenia. Brain Res Mol Brain Res. 2004;132:95–104. doi: 10.1016/j.molbrainres.2004.09.029. [DOI] [PubMed] [Google Scholar]
  • 227.Joober R, Boksa P, Benkelfat C, et al. Genetics of schizophrenia: from animal models to clinical studies. J Psychiatry Neurosci. 2002;27:336–47. [PMC free article] [PubMed] [Google Scholar]
  • 228.Bender HU, Almashanu S, Steel G, et al. Functional consequences of PRODH missense mutations. Am J Hum Genet. 2005;76:409–20. doi: 10.1086/428142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Kempf L, Nicodemus KK, Kolachana B, et al. Functional polymorphisms in PRODH are associated with risk and protection for schizophrenia and fronto-striatal structure and function. PLoS Genet. 2008;4:e1000252. doi: 10.1371/journal.pgen.1000252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Zinkstok J, Schmitz N, van Amelsvoort T, et al. Genetic variation in COMT and PRODH is associated with brain anatomy in patients with schizophrenia. Genes Brain Behav. 2008;7:61–9. doi: 10.1111/j.1601-183X.2007.00326.x. [DOI] [PubMed] [Google Scholar]
  • 231.Liu H, Heath SC, Sobin C, et al. Genetic variation at the 22q11 PRODH2/DGCR6 locus presents an unusual pattern and increases susceptibility to schizophrenia. Proc Natl Acad Sci U S A. 2002;99:3717–22. doi: 10.1073/pnas.042700699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Jacquet H, Raux G, Thibaut F, et al. PRODH mutations and hyperprolinemia in a subset of schizophrenic patients. Hum Mol Genet. 2002;11:2243–9. doi: 10.1093/hmg/11.19.2243. [DOI] [PubMed] [Google Scholar]
  • 233.Li T, Ma X, Sham PC, et al. Evidence for association between novel polymorphisms in the PRODH gene and schizophrenia in a Chinese population. Am J Med Genet B Neuropsychiatr Genet. 2004;129B:13–5. doi: 10.1002/ajmg.b.30049. [DOI] [PubMed] [Google Scholar]
  • 234.Jacquet H, Demily C, Houy E, et al. Hyperprolinemia is a risk factor for schizoaffective disorder. Mol Psychiatry. 2005;10:479–85. doi: 10.1038/sj.mp.4001597. [DOI] [PubMed] [Google Scholar]
  • 235.Ma X, Sun J, Yao J, et al. A quantitative association study between schizotypal traits and COMT, PRODH and BDNF genes in a healthy Chinese population. Psychiatry Res. 2007;153:7–15. doi: 10.1016/j.psychres.2007.02.003. [DOI] [PubMed] [Google Scholar]
  • 236.Ohtsuki T, Tanaka S, Ishiguro H, et al. Failure to find association between PRODH deletion and schizophrenia. Schizophr Res. 2004;67:111–3. doi: 10.1016/s0920-9964(03)00160-9. [DOI] [PubMed] [Google Scholar]
  • 237.Williams HJ, Williams N, Spurlock G, et al. Association between PRODH and schizophrenia is not confirmed. Mol Psychiatry. 2003;8:644–5. doi: 10.1038/sj.mp.4001276. [DOI] [PubMed] [Google Scholar]
  • 238.Roussos P, Giakoumaki SG, Bitsios P. A risk PRODH haplotype affects sensorimotor gating, memory, schizotypy, and anxiety in healthy male subjects. Biol Psychiatry. 2009;65:1063–70. doi: 10.1016/j.biopsych.2009.01.003. [DOI] [PubMed] [Google Scholar]
  • 239.Stark KL, Burt RA, Gogos JA, et al. Analysis of prepulse inhibition in mouse lines overexpressing 22q11.2 orthologues. Int J Neuropsychopharmacol. 2009 Jun 11;:1–7. doi: 10.1017/S1461145709000492. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 240.Gogos JA, Santha M, Takacs Z, et al. The gene encoding proline dehydrogenase modulates sensorimotor gating in mice. Nat Genet. 1999;21:434–9. doi: 10.1038/7777. [DOI] [PubMed] [Google Scholar]
  • 241.Paterlini M, Zakharenko SS, Lai WS, et al. Transcriptional and behavioral interaction between 22q11.2 orthologs modulates schizophrenia-related phenotypes in mice. Nat Neurosci. 2005;8:1586–94. doi: 10.1038/nn1562. [DOI] [PubMed] [Google Scholar]
  • 242.Ross CA, Margolis RL, Reading SA, et al. Neurobiology of schizophrenia. Neuron. 2006;52:139–53. doi: 10.1016/j.neuron.2006.09.015. [DOI] [PubMed] [Google Scholar]
  • 243.Brisch R, Bernstein HG, Krell D, et al. Dopamine-glutamate abnormalities in the frontal cortex associated with the catechol-O-methyltransferase (COMT) in schizophrenia. Brain Res. 2009;1269:166–75. doi: 10.1016/j.brainres.2009.02.039. [DOI] [PubMed] [Google Scholar]
  • 244.Rutherford K, Daggett V. A hotspot of inactivation: the A22S and V108M polymorphisms individually destabilize the active site structure of catechol O-methyltransferase. Biochemistry. 2009;48:6450–60. doi: 10.1021/bi900174v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Honea R, Verchinski BA, Pezawas L, et al. Impact of interacting functional variants in COMT on regional gray matter volume in human brain. Neuroimage. 2009;45:44–51. doi: 10.1016/j.neuroimage.2008.10.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Fan JB, Zhang CS, Gu NF, et al. Catechol-O-methyltransferase gene Val/Met functional polymorphism and risk of schizophrenia: a large-scale association study plus meta-analysis. Biol Psychiatry. 2005;57:139–44. doi: 10.1016/j.biopsych.2004.10.018. [DOI] [PubMed] [Google Scholar]
  • 247.Gupta M, Bhatnagar P, Grover S, et al. Association studies of catechol-O-methyltransferase (COMT) gene with schizophrenia and response to antipsychotic treatment. Pharmacogenomics. 2009;10:385–97. doi: 10.2217/14622416.10.3.385. [DOI] [PubMed] [Google Scholar]
  • 248.Handoko HY, Nyholt DR, Hayward NK, et al. Separate and interacting effects within the catechol-O-methyltransferase (COMT) are associated with schizophrenia. Mol Psychiatry. 2005;10:589–97. doi: 10.1038/sj.mp.4001606. [DOI] [PubMed] [Google Scholar]
  • 249.Okochi T, Ikeda M, Kishi T, et al. Meta-analysis of association between genetic variants in COMT and schizophrenia: an update. Schizophr Res. 2009;110:140–8. doi: 10.1016/j.schres.2009.02.019. [DOI] [PubMed] [Google Scholar]
  • 250.Williams HJ, Glaser B, Williams NM, et al. No association between schizophrenia and polymorphisms in COMT in two large samples. Am J Psychiatry. 2005;162:1736–8. doi: 10.1176/appi.ajp.162.9.1736. [DOI] [PubMed] [Google Scholar]
  • 251.Quednow BB, Wagner M, Mossner R, et al. Sensorimotor gating of schizophrenia patients depends on catechol O-methyltransferase Val158Met polymorphism. Schizophr Bull. 2010;36:341–6. doi: 10.1093/schbul/sbn088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Barnett JH, Jones PB, Robbins TW, et al. Effects of the catechol-O-methyltransferase Val158Met polymorphism on executive function: a meta-analysis of the Wisconsin Card Sort Test in schizophrenia and healthy controls. Mol Psychiatry. 2007;12:502–9. doi: 10.1038/sj.mp.4001973. [DOI] [PubMed] [Google Scholar]
  • 253.Liao SY, Lin SH, Liu CM, et al. Genetic variants in COMT and neurocognitive impairment in families of patients with schizophrenia. Genes Brain Behav. 2009;8:228–37. doi: 10.1111/j.1601-183X.2008.00467.x. [DOI] [PubMed] [Google Scholar]
  • 254.Docherty AR, Sponheim SR. Anhedonia as a phenotype for the Val158Met COMT polymorphism in relatives of patients with schizophrenia. J Abnorm Psychol. 2008;117:788–98. doi: 10.1037/a0013745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Gogos JA, Morgan M, Luine V, et al. Catechol-O-methyltransferase-deficient mice exhibit sexually dimorphic changes in catecholamine levels and behavior. Proc Natl Acad Sci U S A. 1998;95:9991–6. doi: 10.1073/pnas.95.17.9991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Huotari M, Santha M, Lucas LR, et al. Effect of dopamine uptake inhibition on brain catecholamine levels and locomotion in catechol-O-methyltransferase-disrupted mice. J Pharmacol Exp Ther. 2002;303:1309–16. doi: 10.1124/jpet.102.043042. [DOI] [PubMed] [Google Scholar]
  • 257.Schmitt A, Parlapani E, Gruber O, et al. Impact of neuregulin-1 on the pathophysiology of schizophrenia in human post-mortem studies. Eur Arch Psychiatry Clin Neurosci. 2008;258 (Suppl 5):35–9. doi: 10.1007/s00406-008-5019-x. [DOI] [PubMed] [Google Scholar]
  • 258.Hall J, Whalley HC, Job DE, et al. A neuregulin 1 variant associated with abnormal cortical function and psychotic symptoms. Nat Neurosci. 2006;9:1477–8. doi: 10.1038/nn1795. [DOI] [PubMed] [Google Scholar]
  • 259.Mata I, Perez-Iglesias R, Roiz-Santianez R, et al. A neuregulin 1 variant is associated with increased lateral ventricle volume in patients with first-episode schizophrenia. Biol Psychiatry. 2009;65:535–40. doi: 10.1016/j.biopsych.2008.10.020. [DOI] [PubMed] [Google Scholar]
  • 260.Scolnick EM, Petryshen T, Sklar P. Schizophrenia: Do the genetics and neurobiology of neuregulin provide a pathogenesis model? Harv Rev Psychiatry. 2006;14:64–77. doi: 10.1080/10673220600642960. [DOI] [PubMed] [Google Scholar]
  • 261.Bakker SC, Hoogendoorn ML, Selten JP, et al. Neuregulin 1: genetic support for schizophrenia subtypes. Mol Psychiatry. 2004;9:1061–3. doi: 10.1038/sj.mp.4001564. [DOI] [PubMed] [Google Scholar]
  • 262.Prata DP, Breen G, Osborne S, et al. An association study of the neuregulin 1 gene, bipolar affective disorder and psychosis. Psychiatr Genet. 2009;19:113–6. doi: 10.1097/YPG.0b013e32832a4f69. [DOI] [PubMed] [Google Scholar]
  • 263.Stefansson H, Sigurdsson E, Steinthorsdottir V, et al. Neuregulin 1 and susceptibility to schizophrenia. Am J Hum Genet. 2002;71:877–92. doi: 10.1086/342734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Jönsson EG, Saetre P, Vares M, et al. DTNBP1, NRG1, DAOA, DAO and GRM3 polymorphisms and schizophrenia: an association study. Neuropsychobiology. 2009;59:142–50. doi: 10.1159/000218076. [DOI] [PubMed] [Google Scholar]
  • 265.Vilella E, Costas J, Sanjuan J, et al. Association of schizophrenia with DTNBP1 but not with DAO, DAOA, NRG1 and RGS4 nor their genetic interaction. J Psychiatr Res. 2008;42:278–88. doi: 10.1016/j.jpsychires.2007.02.005. [DOI] [PubMed] [Google Scholar]
  • 266.Li D, Collier DA, He L. Meta-analysis shows strong positive association of the neuregulin 1 (NRG1) gene with schizophrenia. Hum Mol Genet. 2006;15:1995–2002. doi: 10.1093/hmg/ddl122. [DOI] [PubMed] [Google Scholar]
  • 267.Gong YG, Wu CN, Xing QH, et al. A two-method meta-analysis of Neuregulin 1(NRG1) association and heterogeneity in schizophrenia. Schizophr Res. 2009;111:109–14. doi: 10.1016/j.schres.2009.03.017. [DOI] [PubMed] [Google Scholar]
  • 268.Rimer M, Barrett DW, Maldonado MA, et al. Neuregulin-1 immunoglobulin-like domain mutant mice: clozapine sensitivity and impaired latent inhibition. Neuroreport. 2005;16:271–5. doi: 10.1097/00001756-200502280-00014. [DOI] [PubMed] [Google Scholar]
  • 269.van den Buuse M, Wischhof L, Lee RX, et al. Neuregulin 1 hypomorphic mutant mice: enhanced baseline locomotor activity but normal psychotropic drug-induced hyperlocomotion and prepulse inhibition regulation. Int J Neuropsychopharmacol. 2009;12:1383–93. doi: 10.1017/S1461145709000388. [DOI] [PubMed] [Google Scholar]
  • 270.Millar JK, Wilson-Annan JC, Anderson S, et al. Disruption of two novel genes by a translocation co-segregating with schizophrenia. Hum Mol Genet. 2000;9:1415–23. doi: 10.1093/hmg/9.9.1415. [DOI] [PubMed] [Google Scholar]
  • 271.Meyer KD, Morris JA. Disc1 regulates granule cell migration in the developing hippocampus. Hum Mol Genet. 2009;18:3286–97. doi: 10.1093/hmg/ddp266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Millar JK, James R, Brandon NJ, et al. Drosoph Inf ServC1 and Drosoph Inf ServC2: discovering and dissecting molecular mechanisms underlying psychiatric illness. Ann Med. 2004;36:367–78. doi: 10.1080/07853890410033603. [DOI] [PubMed] [Google Scholar]
  • 273.Prata DP, Mechelli A, Fu CH, et al. Effect of disrupted-in-schizophrenia-1 on pre-frontal cortical function. Mol Psychiatry. 2008;13:915–7. 909. doi: 10.1038/mp.2008.76. [DOI] [PubMed] [Google Scholar]
  • 274.Szeszko PR, Hodgkinson CA, Robinson DG, et al. Drosoph Inf ServC1 is associated with prefrontal cortical gray matter and positive symptoms in schizophrenia. Biol Psychol. 2008;79:103–10. doi: 10.1016/j.biopsycho.2007.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Takahashi T, Suzuki M, Tsunoda M, et al. The disrupted-in-schizophrenia-1 Ser704Cys polymorphism and brain morphology in schizophrenia. Psychiatry Res. 2009;172:128–35. doi: 10.1016/j.pscychresns.2009.01.005. [DOI] [PubMed] [Google Scholar]
  • 276.Blackwood DH, Fordyce A, Walker MT, et al. Schizophrenia and affective disorders — cosegregation with a translocation at chromosome 1q42 that directly disrupts brain-expressed genes: clinical and P300 findings in a family. Am J Hum Genet. 2001;69:428–33. doi: 10.1086/321969. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Ekelund J, Hovatta I, Parker A, et al. Chromosome 1 loci in Finnish schizophrenia families. Hum Mol Genet. 2001;10:1611–7. doi: 10.1093/hmg/10.15.1611. [DOI] [PubMed] [Google Scholar]
  • 278.Sachs NA, Sawa A, Holmes SE, et al. A frameshift mutation in disrupted in schizophrenia 1 in an American family with schizophrenia and schizoaffective disorder. Mol Psychiatry. 2005;10:758–64. doi: 10.1038/sj.mp.4001667. [DOI] [PubMed] [Google Scholar]
  • 279.Saetre P, Agartz I, De FA, et al. Association between a disrupted-in-schizophrenia 1 (Drosoph Inf ServC1) single nucleotide polymorphism and schizophrenia in a combined Scandinavian case-control sample. Schizophr Res. 2008;106:237–41. doi: 10.1016/j.schres.2008.08.024. [DOI] [PubMed] [Google Scholar]
  • 280.Schumacher J, Laje G, Abou JR, et al. The Drosoph Inf ServC locus and schizophrenia: evidence from an association study in a central European sample and from a meta-analysis across different European populations. Hum Mol Genet. 2009;18:2719–27. doi: 10.1093/hmg/ddp204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Tomppo L, Hennah W, Miettunen J, et al. Association of variants in Drosoph Inf ServC1 with psychosis-related traits in a large population cohort. Arch Gen Psychiatry. 2009;66:134–41. doi: 10.1001/archgenpsychiatry.2008.524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Clapcote SJ, Lipina TV, Millar JK, et al. Behavioral phenotypes of Disc1 missense mutations in mice. Neuron. 2007;54:387–402. doi: 10.1016/j.neuron.2007.04.015. [DOI] [PubMed] [Google Scholar]
  • 283.Kvajo M, McKellar H, Arguello PA, 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;105:7076–81. doi: 10.1073/pnas.0802615105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Pletnikov MV, Ayhan Y, Nikolskaia O, et al. Inducible expression of mutant human Drosoph Inf ServC1 in mice is associated with brain and behavioral abnormalities reminiscent of schizophrenia. Mol Psychiatry. 2008;13:173–86. 115. doi: 10.1038/sj.mp.4002079. [DOI] [PubMed] [Google Scholar]
  • 285.Shen S, Lang B, Nakamoto C, et al. Schizophrenia-related neural and behavioral phenotypes in transgenic mice expressing truncated Disc1. J Neurosci. 2008;28:10893–904. doi: 10.1523/JNEUROSCI.3299-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Yue W, Kang G, Zhang Y, et al. Association of DAOA polymorphisms with schizophrenia and clinical symptoms or therapeutic effects. Neurosci Lett. 2007;416:96–100. doi: 10.1016/j.neulet.2007.01.056. [DOI] [PubMed] [Google Scholar]
  • 287.Chumakov I, Blumenfeld M, Guerassimenko O, et al. Genetic and physiological data implicating the new human gene G72 and the gene for D-amino acid oxidase in schizophrenia. Proc Natl Acad Sci U S A. 2002;99:13675–80. doi: 10.1073/pnas.182412499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Shinkai T, DeLuac V, Hwang R, et al. Association analyses of the DAOA/G30 and D-amino-acid oxidase genes in schizophrenia: further evidence for a role in schizophrenia. Neuromolecular Med. 2007;9:169–77. doi: 10.1007/BF02685890. [DOI] [PubMed] [Google Scholar]
  • 289.Shin HD, Park BL, Kim EM, et al. Association analysis of G72/G30 polymorphisms with schizophrenia in the Korean population. Schizophr Res. 2007;96:119–24. doi: 10.1016/j.schres.2007.05.004. [DOI] [PubMed] [Google Scholar]
  • 290.Shi J, Badner JA, Gershon ES, et al. Further evidence for an association of G72/G30 with schizophrenia in Chinese. Schizophr Res. 2009;107:324–6. doi: 10.1016/j.schres.2008.07.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Liu YL, Fann CS, Liu CM, et al. No association of G72 and D-amino acid oxidase genes with schizophrenia. Schizophr Res. 2006;87:15–20. doi: 10.1016/j.schres.2006.06.020. [DOI] [PubMed] [Google Scholar]
  • 292.Li D, He L. G72/G30 genes and schizophrenia: a systematic meta-analysis of association studies. Genetics. 2007;175:917–22. doi: 10.1534/genetics.106.061796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Jansen A, Krach S, Krug A, et al. A putative high risk diplotype of the G72 gene is in healthy individuals associated with better performance in working memory functions and altered brain activity in the medial temporal lobe. Neuroimage. 2009;45:1002–8. doi: 10.1016/j.neuroimage.2008.12.054. [DOI] [PubMed] [Google Scholar]
  • 294.Otte DM, Bilkei-Gorzo A, Filiou MD, et al. Behavioral changes in G72/G30 transgenic mice. Eur Neuropsychopharmacol. 2009;19:339–48. doi: 10.1016/j.euroneuro.2008.12.009. [DOI] [PubMed] [Google Scholar]
  • 295.Owen MJ, Williams NM, O’Donovan MC. Dysbindin-1 and schizophrenia: from genetics to neuropathology. J Clin Invest. 2004;113:1255–7. doi: 10.1172/JCI21470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Weickert CS, Straub RE, McClintock BW, et al. Human dysbindin (DTNBP1) gene expression in normal brain and in schizophrenic prefrontal cortex and midbrain. Arch Gen Psychiatry. 2004;61:544–55. doi: 10.1001/archpsyc.61.6.544. [DOI] [PubMed] [Google Scholar]
  • 297.Duan J, Martinez M, Sanders AR, et al. DTNBP1 (Dystrobrevin binding protein 1) and schizophrenia: association evidence in the 3′ end of the gene. Hum Hered. 2007;64:97–106. doi: 10.1159/000101961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Pae CU, Mandelli L, De RD, et al. Dysbindin gene (DTNBP1) and schizophrenia in Korean population. Eur Arch Psychiatry Clin Neurosci. 2009;259:137–42. doi: 10.1007/s00406-008-0830-y. [DOI] [PubMed] [Google Scholar]
  • 299.Tosato S, Ruggeri M, Bonetto C, et al. Association study of dysbindin gene with clinical and outcome measures in a representative cohort of Italian schizophrenic patients. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:647–59. doi: 10.1002/ajmg.b.30484. [DOI] [PubMed] [Google Scholar]
  • 300.Datta SR, McQuillin A, Puri V, et al. Failure to confirm allelic and haplotypic association between markers at the chromosome 6p22.3 dystrobrevin-binding protein 1 (DTNBP1) locus and schizophrenia. Behav Brain Funct. 2007;3:50. doi: 10.1186/1744-9081-3-50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Holliday EG, Handoko HY, James MR, et al. Association study of the dystrobrevin-binding gene with schizophrenia in Australian and Indian samples. Twin Res Hum Genet. 2006;9:531–9. doi: 10.1375/183242706778025035. [DOI] [PubMed] [Google Scholar]
  • 302.Peters K, Wiltshire S, Henders AK, et al. Comprehensive analysis of tagging sequence variants in DTNBP1 shows no association with schizophrenia or with its composite neurocognitive endophenotypes. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:1159–66. doi: 10.1002/ajmg.b.30741. [DOI] [PubMed] [Google Scholar]
  • 303.Li D, He L. Association study between the dystrobrevin binding protein 1 gene (DTNBP1) and schizophrenia: a meta-analysis. Schizophr Res. 2007;96:112–8. doi: 10.1016/j.schres.2007.05.017. [DOI] [PubMed] [Google Scholar]
  • 304.Donohoe G, Morris DW, Clarke S, et al. Variance in neurocognitive performance is associated with dysbindin-1 in schizophrenia: a preliminary study. Neuropsychologia. 2007;45:454–8. doi: 10.1016/j.neuropsychologia.2006.06.016. [DOI] [PubMed] [Google Scholar]
  • 305.Hashimoto R, Noguchi H, Hori H, et al. A genetic variation in the dysbindin gene (DTNBP1) is associated with memory performance in healthy controls. World J Biol Psychiatry. 2010;11:431–8. doi: 10.1080/15622970902736503. [DOI] [PubMed] [Google Scholar]
  • 306.Takao K, Toyama K, Nakanishi K, et al. Impaired long-term memory retention and working memory in sdy mutant mice with a deletion in Dtnbp1, a susceptibility gene for schizophrenia. Mol Brain. 2008;1:11. doi: 10.1186/1756-6606-1-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Cox MM, Tucker AM, Tang J, et al. Neurobehavioral abnormalities in the dysbindin-1 mutant, sandy, on a C57BL/6J genetic background. Genes Brain Behav. 2009;8:390–7. doi: 10.1111/j.1601-183X.2009.00477.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Feng YQ, Zhou ZY, He X, et al. Dysbindin deficiency in sandy mice causes reduction of snapin and displays behaviors related to schizophrenia. Schizophr Res. 2008;106:218–28. doi: 10.1016/j.schres.2008.07.018. [DOI] [PubMed] [Google Scholar]
  • 309.Hattori S, Murotani T, Matsuzaki S, et al. Behavioral abnormalities and dopamine reductions in sdy mutant mice with a deletion in Dtnbp1, a susceptibility gene for schizophrenia. Biochem Biophys Res Commun. 2008;373:298–302. doi: 10.1016/j.bbrc.2008.06.016. [DOI] [PubMed] [Google Scholar]
  • 310.Sinibaldi L, De LA, Bellacchio E, et al. Mutations of the Nogo-66 receptor (RTN4R) gene in schizophrenia. Hum Mutat. 2004;24:534–5. doi: 10.1002/humu.9292. [DOI] [PubMed] [Google Scholar]
  • 311.Novak G, Kim D, Seeman P, et al. Schizophrenia and Nogo: elevated mRNA in cortex, and high prevalence of a homozygous CAA insert. Brain Res Mol Brain Res. 2002;107:183–9. doi: 10.1016/s0169-328x(02)00492-8. [DOI] [PubMed] [Google Scholar]
  • 312.Novak G, Tallerico T, Nogo A. B and C expression in schizophrenia, depression and bipolar frontal cortex, and correlation of Nogo expression with CAA/TATC polymorphism in 3′-UTR. Brain Res. 2006;1120:161–71. doi: 10.1016/j.brainres.2006.08.071. [DOI] [PubMed] [Google Scholar]
  • 313.Chen W, Gu N, Duan S, et al. No association between the genetic polymorphisms within RTN4 and schizophrenia in the Chinese population. Neurosci Lett. 2004;365:23–7. doi: 10.1016/j.neulet.2004.04.021. [DOI] [PubMed] [Google Scholar]
  • 314.Gregório SP, Mury FB, Ojopi EB, et al. Nogo CAA 3′UTR insertion polymorphism is not associated with schizophrenia nor with bipolar disorder. Schizophr Res. 2005;75:5–9. doi: 10.1016/j.schres.2004.11.010. [DOI] [PubMed] [Google Scholar]
  • 315.Budel S, Padukkavidana T, Liu BP, et al. Genetic variants of Nogo-66 receptor with possible association to schizophrenia block myelin inhibition of axon growth. J Neurosci. 2008;28:13161–72. doi: 10.1523/JNEUROSCI.3828-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Hsu R, Woodroffe A, Lai WS, et al. Nogo Receptor 1 (RTN4R) as a candidate gene for schizophrenia: analysis using human and mouse genetic approaches. PLoS ONE. 2007;2:e1234. doi: 10.1371/journal.pone.0001234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 317.Meng J, Shi Y, Zhao X, et al. No association between the genetic polymorphisms in the RTN4R gene and schizophrenia in the Chinese population. J Neural Transm. 2007;114:249–54. doi: 10.1007/s00702-006-0538-y. [DOI] [PubMed] [Google Scholar]
  • 318.Buckholtz JW, Meyer-Lindenberg A, Honea RA, et al. Allelic variation in RGS4 impacts functional and structural connectivity in the human brain. J Neurosci. 2007;27:1584–93. doi: 10.1523/JNEUROSCI.5112-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Chowdari KV, Mirnics K, Semwal P, et al. Association and linkage analyses of RGS4 polymorphisms in schizophrenia. Hum Mol Genet. 2002;11:1373–80. doi: 10.1093/hmg/11.12.1373. [DOI] [PubMed] [Google Scholar]
  • 320.So HC, Chen RY, Chen EY, et al. An association study of RGS4 polymorphisms with clinical phenotypes of schizophrenia in a Chinese population. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:77–85. doi: 10.1002/ajmg.b.30577. [DOI] [PubMed] [Google Scholar]
  • 321.Talkowski ME, Seltman H, Bassett AS, et al. Evaluation of a susceptibility gene for schizophrenia: genotype based meta-analysis of RGS4 polymorphisms from thirteen independent samples. Biol Psychiatry. 2006;60:152–62. doi: 10.1016/j.biopsych.2006.02.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Guo S, Tang W, Shi Y, et al. RGS4 polymorphisms and risk of schizophrenia: an association study in Han Chinese plus meta-analysis. Neurosci Lett. 2006;406:122–7. doi: 10.1016/j.neulet.2006.07.028. [DOI] [PubMed] [Google Scholar]
  • 323.Li D, He L. Association study of the G-protein signaling 4 (RGS4) and proline dehydrogenase (PRODH) genes with schizophrenia: a meta-analysis. Eur J Hum Genet. 2006;14:1130–5. doi: 10.1038/sj.ejhg.5201680. [DOI] [PubMed] [Google Scholar]
  • 324.Grillet N, Pattyn A, Contet C, et al. Generation and characterization of Rgs4 mutant mice. Mol Cell Biol. 2005;25:4221–8. doi: 10.1128/MCB.25.10.4221-4228.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Mukai J, Liu H, Burt RA, et al. Evidence that the gene encoding ZDHHC8 contributes to the risk of schizophrenia. Nat Genet. 2004;36:725–31. doi: 10.1038/ng1375. [DOI] [PubMed] [Google Scholar]
  • 326.Liu H, Abecasis GR, Heath SC, et al. Genetic variation in the 22q11 locus and susceptibility to schizophrenia. Proc Natl Acad Sci U S A. 2002;99:16859–64. doi: 10.1073/pnas.232186099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Chen WY, Shi YY, Zheng YL, et al. Case-control study and transmission disequilibrium test provide consistent evidence for association between schizophrenia and genetic variation in the 22q11 gene ZDHHC8. Hum Mol Genet. 2004;13:2991–5. doi: 10.1093/hmg/ddh322. [DOI] [PubMed] [Google Scholar]
  • 328.Faul T, Gawlik M, Bauer M, et al. ZDHHC8 as a candidate gene for schizophrenia: analysis of a putative functional intronic marker in case-control and family-based association studies. BMC Psychiatry. 2005;5:35. doi: 10.1186/1471-244X-5-35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 329.Otani K, Ujike H, Tanaka Y, et al. The ZDHHC8 gene did not associate with bipolar disorder or schizophrenia. Neurosci Lett. 2005;390:166–70. doi: 10.1016/j.neulet.2005.08.019. [DOI] [PubMed] [Google Scholar]
  • 330.Mukai J, Dhilla A, Drew LJ, et al. Palmitoylation-dependent neurodevelopmental deficits in a mouse model of 22q11 microdeletion. Nat Neurosci. 2008;11:1302–10. doi: 10.1038/nn.2204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Nicodemus KK, Kolachana BS, Vakkalanka R, et al. Evidence for statistical epistasis between catechol-O-methyltransferase (COMT) and polymorphisms in RGS4, G72 (DAOA), GRM3, and Drosoph Inf ServC1: influence on risk of schizophrenia. Hum Genet. 2007;120:889–906. doi: 10.1007/s00439-006-0257-3. [DOI] [PubMed] [Google Scholar]
  • 332.Moncrieff J, Cohen D. How do psychiatric drugs work? BMJ. 2009;338:b1963. doi: 10.1136/bmj.b1963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Lesch KP. Gene transfer to the brain: Emerging therapeutic strategy in psychiatry? Biol Psychiatry. 1999;45:247–53. doi: 10.1016/s0006-3223(98)00236-4. [DOI] [PubMed] [Google Scholar]
  • 334.Clapcote SJ, Roder JC. Deletion polymorphism of Disc1 is common to all 129 mouse substrains: implications for gene-targeting studies of brain function. Genetics. 2006;173:2407–10. doi: 10.1534/genetics.106.060749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Anagnostopoulos AV, Mobraaten LE, Sharp JJ, et al. Transgenic and knockout databases: behavioral profiles of mouse mutants. Physiol Behav. 2001;73:675–89. doi: 10.1016/s0031-9384(01)00525-x. [DOI] [PubMed] [Google Scholar]
  • 336.Clapcote SJ, Lazar NL, Bechard AR, et al. NIH Swiss and Black Swiss mice have retinal degeneration and performance deficits in cognitive tests. Comp Med. 2005;55:310–6. [PubMed] [Google Scholar]
  • 337.Clapcote SJ, Lazar NL, Bechard AR, et al. Effects of the rd1 mutation and host strain on hippocampal learning in mice. Behav Genet. 2005;35:591–601. doi: 10.1007/s10519-005-5634-5. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Psychiatry & Neuroscience : JPN are provided here courtesy of Canadian Medical Association

RESOURCES