Skip to main content
Neural Plasticity logoLink to Neural Plasticity
. 2012 Oct 24;2012:832757. doi: 10.1155/2012/832757

Effects of Antipsychotics on Dentate Gyrus Stem Cell Proliferation and Survival in Animal Models: A Critical Update

Gerburg Keilhoff 1,*, Paolo Fusar-Poli 2, Axel Becker 3
PMCID: PMC3488410  PMID: 23150836

Abstract

Schizophrenia is a complex psychiatric disorder. Although a number of different hypotheses have been developed to explain its aetiopathogenesis, we are far from understanding it. There is clinical and experimental evidence indicating that neurodevelopmental factors play a major role. Disturbances in neurodevelopment might result in alterations of neuroanatomy and neurochemistry, leading to the typical symptoms observed in schizophrenia. The present paper will critically address the neurodevelopmental models underlying schizophrenia by discussing the effects of typical and atypical antipsychotics in animal models. We will specifically discuss the vitamin D deficiency model, the poly I:C model, the ketamine model, and the postnatal ventral hippocampal lesion model, all of which reflect core neurodevelopmental issues underlying schizophrenia onset.

1. Introduction

Schizophrenia is a complex psychiatric disorder which is characterized by a defined set of symptoms usually grouped into positive symptoms, negative symptoms, cognitive impairment, psychosocial impairments, and poor quality of life. It is commonly described as a developmental disorder, with onset in the early adulthood or adolescence and involving several genetic and environmental factors. The causes of schizophrenia are unknown. However, several hypotheses have been tested in the recent research. One of the most accepted theories is the “two hit hypothesis.” Such a hypothesis proposes that an early disturbance is necessary but not sufficient to cause an increased vulnerability to schizophrenia. Thus, an early neurodevelopmental insult is requested to interact with either normal or abnormal postpubertal brain maturation to fully produce late neurodevelopmental brain structural and functional changes [14].

Evidence indicating a neurodevelopmental origin of schizophrenia is grounded on extensive research performed over the past two decades. In particular, a novel approach has allowed clinicians to specifically investigate the early phases of psychosis and to clarify the mechanisms underlying the onset of the illness. This approach has been variably termed as ultrahigh risk, at risk mental state, or clinical high risk [5]. This putatively prodromal psychotic phase is associated with an enhanced risk of developing the illness as compared to the general population (1%), ranging from 18% at six months up to 36% after three years [6]. The majority (73%) of the individuals developing a psychotic illness will transit towards a schizophrenia spectrum disorder [7]. The clinical high risk state for psychosis is also characterized by significant cognitive impairments [8] and deficits in social functioning and quality of life [5]. These alterations are associated with underlying neurodevelopmental abnormalities in the structure [911], function [1214], connectivity [15], and neurochemistry [1618] of the brain, resembling those observed in the established phase of the illness [19]. Interest in this area has exponentially grown to the extent that a new diagnostic category is being discussed in the forthcoming DSM-5 [20]. These findings taken together indicated that schizophrenia is characterized by dynamic neurobiological changes since its earliest phases. In theory, the early phases of schizophrenia can thus be particularly amenable to treatments that can impact the underlying neurobiology, including antipsychotics. The present paper will critically address this point, focusing on the role played by the effect of antipsychotics on the neurogenesis during the onset of schizophrenia. These issues will be discussed in the light of the recent advances in animal models.

2. Methodological Approach

In the following sections, we will provide an update on the effects of antipsychotics on cell proliferation in animal models used in schizophrenia research. This critical paper is limited to models most traditionally employed in the laboratories, whereby all models reflect neurodevelopmental aspects. In Table 1, a selection of most relevant publications in the last 13 years is given. A survey of effects of neuroleptics on hippocampal neurogenesis is provided in Table 2.

Table 1.

Critical selection of most relevant papers published in the last 13 years.

Author Year Finding Reason for selection
Bayer et al. [1] 1999 Genetic or environmental first hit affects brain development, a second hit later in life initiated the outbreak of schizophrenia Comprehensive hypothesis concerning schizophrenia etiology

Arnold and Watt [21] 2001 Number of immature cells is increased in the olfactory epithelium of schizophrenics First report on altered cell density in schizophrenics

Reif et al. [22] 2006 Cell proliferation is diminished in the dentate gyrus of schizophrenics First report on altered cell proliferation in the human brain

Kippin et al. [23]
Wakade et al. [24]
Wang et al. [25]
2005
2002
2004
Typical and atypical neuroleptics enhance neurogenesis in the subventricular zone Effects of neuroleptics on neurogenesis in the subventricular zone

Kodama et al. [26]
Wakade et al. [24]
Halim et al. [27]
2004
2002
2004
Atypical but not typical neuroleptics interfere with hippocampal neurogenesis Effects of neuroleptics on neurogenesis in the hippocampus

Piontkewitz et al. [28] 2012 Risperidone partially restored impaired neurogenesis in poly I:C offspring Effect of an atypical neuroleptic on neurogenesis in an model of maternal infection

Keilhoff et al. [29] 2010 Subchronic treatment with haloperidol ameliorated decreased neurogenesis and normalised behaviour in vitamin D-deficient rats Effect of neuroleptics on neurogenesis in the vitamin D model

Keilhoff et al. [30] 2010 Risperidone and haloperidol promoted survival in stem cells in the hippocampus of rats subchronically treated with ketamine Effects of typical and atypical neuroleptics on neurogenesis in the ketamine model

Table 2.

Survey of the effects of neuroleptics on hippocampal neurogenesis.

Model Effect on neuroleptics on neurogenesis in the hippocampus Literature
Vitamin D deficiency Haloperidol ↑ [29]
Maternal infection Risperidone ↑ [28]
Clozapine Ø [64]
Ketamine Haloperidol ↑* [30]
Risperidone ↑* [30]
Phencyclidine Risperidone Ø [65]
Clozapine ↑ [66]
Postnatal lesion of
ventral hippocampus
Haloperidol Ø [67], indirect

Ø: no effect, ↑: increase,  ↗: increase after application of low doses, *: cell survival.

3. Modelling Schizophrenia with Animal Paradigms

The development of animal models is a crucial issue in biological psychiatry for the study of alterations in neurochemistry, neuroanatomy, and behaviours resembling those observed in schizophrenia. Similarly, it can be useful for the discovery and development of effective treatments such as antipsychotic molecules. However, heterogeneity of the clinical symptoms of schizophrenia and the incomplete knowledge about the cause and progression of the illness make the development of valid animal models particularly difficult. Moreover, there is scepticism as to what extent the behaviour of animals can actually reflect highly complex disorders such as schizophrenia. Since each animal model is developed to target only each specific domain of schizophrenia, different complementary approaches are usually needed. Therefore (i) developmental, (ii) genetic, and (iii) pharmacological models have been used in experimental schizophrenia research [3138].

(i) The developmental hypothesis and the respective animal models proceed from the assumption that malformations at very early stages of neurogenesis result in structural abnormalities of the adult brain [3941]. Under this scenario, the pathogenesis of schizophrenia onset is attributed to abnormal neuronal development and/or re-organization of neuronal circuits in the frontal cortex or in limbic structures. Hippocampal volume reductions [42, 43], hippocampal shape deformation, or abnormalities in the hippocampal cell density [44] have been consistently reported. Many of these alterations, however, are essentially based on genetic deficits impacting the brain development (see below).

(ii) Genetic animal models are developed by translation of human genetic mutations into animals (for review: [33, 4547]). They include whole-body mutant mouse strains ([48] (Reelin); [49] (Neuregulin-1)), mutant mouse strains in which distinct genes have been knocked down in a tissue- or cell-type-specific manner ([50] (NRG-1/ErbB); [51] (DISC)), and transgenic mice that overexpress schizophrenia-relevant genes ([52] (dopamine D2 receptor); [53] (Neuregulin-1)). More advanced and complex models, however, are under development. These models combine several environmental and/or genetic factors to better account for the complex aetiology of schizophrenia [54]. For example, combined genetic disruption of the NMDA receptor subunit 1 [5558], the dopamine D2 and D4 receptors, the dopamine transporter [5961], and, the mutants in the dopamine-degrading enzyme catechol-O-methyl-transferase [62, 63] may provide a unique tool to study imbalance in the functional regulation of neurotransmitters implicated in schizophrenia.

(iii) Finally, the development of schizophrenia-relevant animal models can also target the pharmacodynamics of common antipsychotic drugs, to identify their molecular substrates, and to optimise their pharmacokinetics, to develop new drugs, or to test potential antipsychotics.

4. Concept of Neurogenesis in Schizophrenia

Neuronal stem cells (NSCs) belong to the class of adult stem cells. They are multipotent and able to generate (only) the specific cell lineages of the nervous system: neurons, astrocytes, and oligodendrocytes [68]. NSCs will be generated throughout the whole life, but with declining intensity. NSCs are primarily located in the subgranular zone (SGZ) of the hippocampal dentate gyrus and in the subventricular zone (SVZ) of the lateral ventricles. But there is also evidence that NSCs are present in multiple areas of the adult brain [69]. Under the influence of their local microenvironment, that is, their niche, NSCs take different developmental pathways/roads of life. NSCs in the SVZ become neuroblasts, migrate towards the rostral migratory stream into the olfactory bulb, and develop into interneurons. NSCs in the SGZ, on the other hand, develop into local dentate granule cells [70].

Even if most evidences were acquired from different animal models, there is converging consensus that adult neurogenesis seems to be essential for different processes, such as learning and memory [71, 72], mood regulation [73, 74], physiological (maintenance) neuroregeneration, neurorestoration after mechanical brain injuries [75, 76], stroke [76, 77], multiple sclerosis [78, 79], and Parkinson's disease [80]. Thus it is not surprising that disturbances of adult neurogenesis are investigated in a wide range of pathological processes including neurodegenerative diseases, brain tumours, seizures, and mental illnesses such as schizophrenia, major depression, dementia, and alcoholism [70].

Because of these reasons, over the past years there has been a growing interest into neurogenesis-relevant research on postmortem human tissues of schizophrenic patients. Arnold and Watt [21] found abnormal neuronal densities in the olfactory epithelium of schizophrenics. Rioux and Arnold [81] demonstrated a deregulated expression of retinoid receptors in schizophrenia and that retinoid signalling plays a central role in neurogenesis. Reif and coworkers [22] were able to demonstrate that the first step of adult neurogenesis, that is, cell proliferation, is diminished in the dentate gyrus of patients suffering from schizophrenia. Atz and coworkers [82] found a genotypic association of NCAM1 polymorphisms with schizophrenia. NCAM expression is a characteristic feature of the postnatal neurogenic niche [83].

Beside differentiation into neurons, adult neuronal stem cells may also undergo gliogenesis. There is an increasing body of evidence for glial pathology in schizophrenia [84]. Oligodendrocyte and myelin dysfunction are perturbed in schizophrenia since its earliest phases. Through changes in synaptic formation and/or function, they can induce cognitive dysfunction, one of the core symptoms of schizophrenia [85, 86]. The activation of astrocytes has been discussed as an important pathogenic factor for the development of schizophrenia [87], and also microglia has been shown to remodel the CNS during development as well as after injuries [8892].

5. Neurogenic Potential of Antipsychotics

Basically, antipsychotic drugs are divided into typical (first-generation antipsychotics) and atypical (second-generation antipsychotics). Within the commonly used typical antipsychotics in early psychosis are butyrophenone (e.g., haloperidol) and phenothiazine (e.g., chlorpromazine) derivates and within the atypical are clozapine, olanzapine, quetiapine, risperidone (reviewed in [93]).

Antipsychotics interfere with neuronal remodelling. Thereby dopaminergic effects seem to be involved [23, 94]. It was suggested that stimulation of dopamine D2 receptors inhibits the proliferation of neuronal stem cells and that tonic endogenous dopamine inhibits their proliferation [88]. Moreover, it was reported that blockade of D2 receptors activates transcription factors which regulate the expression of genes of neuronal growth factors [95].

Olanzapine, which has less affinity for the dopaminergic receptors [96], is also able to enhance SVZ [24, 25] and hippocampal neurogenesis [26]. The latter, on the other hand, is not influenced by haloperidol [24, 27] clearly indicating that typical and atypical antipsychotics differentially regulate neurogenesis. In support to this notion, it was also reported that typical and atypical antipsychotics differentially induce neuronal plasticity and synaptic remodelling. Atypical but not typical antipsychotics are effective not only in the striatum but also in the prefrontal cortex and hippocampus [94]. With respect to the schizophrenia-relevant morphometric changes and cellular abnormalities in the grey matter, the reported effects of antipsychotics in the prefrontal cortex are particularly interesting (reviewed in [97]). Interestingly, long-term antipsychotic treatment induced glial but not neuronal cell proliferation in monkeys and that with no difference between typical and atypical antipsychotics [98]. All these results, however, do not challenge the scientific consensus that the adult cortex under physiological conditions belongs to the so-called nonneurogenic tissue.

6. Mechanisms Underlying the Effects of Antipsychotics on Neurogenesis

There are different possible mechanisms by which antipsychotics realize their influence on cell proliferation/neurogenesis. Earlier studies indicate that a couple of factors, for example, trophic and transcription factors, can interact in a fine-tuned network. Clozapine, for example, selectively increased FGF-2 (fibroblast growth factor-2, belongs to trophic factors) in the striatum [99]. In the hippocampus, FGF-2 is induced by quetiapine, but only when the NMDA receptor system is downregulated [100]. Other studies addressed the schizophrenia-relevant role of BDNF (brain-derived neurotrophic factor). However, the findings on BDNF status in naïve patients as well as in patients treated with antipsychotics are highly discrepant (reviewed by [101105]). In animals, BDNF in the hippocampus was decreased by haloperidol and high-dosed risperidone [106108], while olanzapine therapy enhanced BDNF [109]. It was also shown that haloperidol reduced NGF (neuronal growth factor) while olanzapine raised NGF levels and risperidone was ineffective on NGF [110, 111]. VEGF (vascular endothelial growth factor, angiogenic neurotrophin) seemed also to be involved in the action of antipsychotics. Haloperidol and olanzapine increased its hippocampal levels [112].

Antipsychotics can additionally influence cell proliferation/neurogenesis via targeting transcription factors implicated in mitotic activity regulation. Thus, it was shown that haloperidol, risperidone, and clozapine affected phosphorylation of extracellular signal-regulated kinases (ERKs) and cyclic adenosine 3′,5′-monophosphate (cAMP) response element (CRE) binding protein (CREB), each with different profiles. In fact, haloperidol and risperidone promoted phosphorylation [113, 114], while clozapine reduced ERK1/2 and CREB phosphorylation [113]. Furthermore, haloperidol treatment of mice increased phosphorylation of Akt1. With respect to the Akt/GSK-3 system, clozapine had similar effects as haloperidol increasing the Akt1 phosphorylation [115].

It is important to note that a direct antipsychotic drug-gene interaction should be taken into close consideration, even if a direct intervention on genes (belonging to the glutamate/NMDA receptor family) has been shown only for haloperidol [116, 117]. Since modulation of progenitor cell proliferation as well as neurogenesis resulting in NMDA receptor modulation has been described [118], these findings set one possible agenda by which a direct antipsychotic drug-gene interaction can become neurogenic.

Assuming that the pathology of the schizophrenia [119, 120] and the cell proliferation/neurogenesis [70, 121] are subjected to epigenetic control mechanisms, future research is needed to address the exact neurogenic mechanisms of antipsychotics adjusting by epigenetic factors.

A comprehensive summary of questions concerning neurogenic actions of antipsychotic drugs is given by Newton and Duman [122].

7. From the Bench: Interplay of Schizophrenia, Neuroleptics, and Neurogenesis

Several pathophysiological models have been proposed to explain schizophrenia and may appear to reflect distinct aspects of this disease. None of the pharmacological, genetic, and neurodevelopmental models have been evaluated in detail for translational relevance or to satisfy requirements of the different levels of validity (face, construct, and predictive validity; for review see [123]). Pharmacological models focused on alterations in the dopaminergic, glutamatergic, serotonergic, and GABAergic neurotransmitter systems [54, 124, 125]. They are based on alterations in these neurotransmitter systems, mimicking the in vivo conditions which are clinically relevant for schizophrenia. These alterations may be manipulated by drug challenges.

8. First Lesson: The Aspect of Maternal Vitamin D Deficiency

Given the apparent polygenic nature of schizophrenia and the limited translational significance of the available pharmacological models, neurodevelopmental models may offer a better chance of success [126]. Different animal models in schizophrenia research were developed to shed light on the developmental aspects of the disease.

In the course of ontogenesis there are two critical phases when the organism is susceptible to disturbances which can contribute to schizophrenia, that is, the embryonic/postdelivery phase (first hit) and during puberty (second hit). The vitamin D deficiency model appears to be useful to study the impact of brain disturbances during embryonic/fetal development [127, 128]. Interestingly, some of the vitamin D deficiency effects could be related to later gestational periods thus possibly expanding the hazardous time window for the neurobiological development of schizophrenia [129]. Moreover, it was shown that the neurosteroid can impact brain development by affecting migration and survival of developing neurones in the brain, by influencing brain levels of neurotrophins and their receptors [130], by altering brain apoptotic activity [131], and by exerting immunoregulatory and neuroprotective effects (for review see [127]).

The vitamin D deficiency model is one of the most commonly explored and used also in our lab [29, 132]. Generally, animals from a normal diet control group which were left untreated revealed a basal level of 5-bromo-2′-deoxyuridine (BrdU) immunolabelling in the hippocampal subgranular zone which was in line with previous reports [133]. Cell typing 5 days after BrdU application offered a large pool of BrdU-labelled cells costained with DCX, a marker for immature neurons (~55%). A colabelling with NeuN, a marker for mature neurons, was only rarely found. The second largest group of BrdU-positive cells (~10%) were round or oval, medium-sized, and immunopositive for nestin, an intermediate filament that is expressed in neuronal stem or progenitor cells, identifying these cells as granule cell precursors. The remaining BrdU-positive cells expressed the common astroglia marker GFAP (~4%), the NG2 proteoglycan, a marker for oligodendroglial precursor cells and/or synantocytes (~4%), or were free of any co-labelling (~7%). There were no obvious differences in the distribution of BrdU-immunoreactive cells at different longitudinal levels of the dentate gyrus.

When cell typing was done 3 weeks after the last BrdU application, about 75% of the BrdU-labelled cells could be identified as granule cells. They had a small round soma, were immunopositive for NeuN, and some of them were shifted from the subgranular cell layer towards the middle part of the granule cell layer. Now, DCX co-labelling was found to be poor. Saline treatment did not alter this BrdU-labelling/costaining pattern.

Prenatal vitamin D deficiency reduced cell proliferation in the subgranular zone. The loss was proportionally distributed between the different cell types. Due to the very little counts of BrdU-positive cells marked with all nonneuronal markers, only the loss of DCX-positive cells was numerically evident.

In control animals, the typical (first-generation) neuroleptic haloperidol significantly increased the total number of BrdU-labelled cells. In vitamin D-deficient mice, the deficiency-induced reduction of cell proliferation was completely normalized by haloperidol resulting in a mitotic activity adequate to the untreated control level. In both cases, haloperidol treatment revealed tendentially more DCX-expressing cells. Moreover, a cytoskeletal hypertrophy of radial glia-like GFAP-positive astrocytes, possibly serving as climbing frame for the migrating neuronal newcomers, was found.

There is general consensus that neuroleptic drugs improve the psychopathology of schizophrenia. Treatment with typical neuroleptics is considered to result in minimal improvement or in worsening of cognitive processes [134], but there are also reports showing that typical neuroleptics provide modest gains in multiple cognitive domains [135]. As hippocampal neurogenesis plays an important role in learning and memory processes [74], we speculate that the previously demonstrated normalization of a vitamin D deficiency-induced habituation deficit in the hole board by haloperidol [136] could result from an at least partially restored mitotic activity. This idea is supported by findings that vitamin D depletion depressed promitotic genes [131, 137]. It is also plausible that there is some kind of exhaustion of the mitotic cell potency due to the overshooting activity in young animals, as it was demonstrated for the vitamin D deficiency model [131] and the NOS knock-out model [138]. Using learning paradigms dependent on hippocampal integrity in subsequent experiments, effects of APDs on both learning behaviour and neurogenesis should be studied in detail.

9. Second Lesson: The Aspect of Maternal Infection

Epidemiological studies have shown that maternal infection and inflammation in definite periods of pregnancy are significantly associated with an increased risk of schizophrenia in the offsprings. Infection with influenza virus [139] or application of polyriboinosinic-polyribocytidylic acid (poly I:C), an inflammatory agent which mimics inflammation by stimulation cytokine release through Toll-like receptor TLR3 activation is accepted models in schizophrenia research (for review see [140]). Prenatal immune stimulation reduces hippocampal neurogenesis [141144]. The beneficial effects of atypical neuroleptic drugs (APDs), on the other hand, have been attributed to their capacity to increase neurogenesis [22, 145147]. Together with the group of Weiner and Piontkewitz [28], some of the present authors studied the effects of adolescent poly I:C and risperidone treatment by analyzing a battery of cellular markers referring to cell proliferation and differentiation of hippocampal cell populations. The offspring of poly I:C-treated dams were characterized by an impaired neurogenesis including a decrease of calretinin-positive neurons, disturbed microvascularization and granular cell density in the dentate area, and a reduction of parvalbumin-expressing interneurons, whose deficit is a well-replicated neuropathological finding in schizophrenia [148]. Risperidone normalized the disturbed cell proliferation and/or survival, the number of calretinin and parvalbumin-expressing cells, and counteracted the disturbance in angiogenesis.

Together with previous reports on deficient hippocampal neurogenesis in offspring of poly I:C-exposed mice [64, 144] and LPS-exposed rats [141], our findings confirm the hypothesis that impaired neurogenesis is an important aetiopathological factor for hippocampal abnormalities and related cognitive dysfunctions in animal models and in patients with schizophrenia [22, 30, 74, 123, 141, 145147, 149, 150]. Studies concerning an influence of antipsychotics in the poly I:C model are rare. Thus, Meyer et al. [64] demonstrated that chronic clozapine treatment had no effect on poly I:C-hampered neurogenesis.

With respect to neurogenesis, our findings about a risperidone-mediated normalization of the byprenatal poly I:C disturbed angiogenesis are of special interest. In general, angiogenesis and neurogenesis are closely linked with each other [151]. Thus, VEGF modulates neurogenesis directly and also subsequently releasing neurotrophic factors such as BDNF [145, 152, 153]. Moreover, in a previous study we showed that administration of risperidone was able to increase VEGF expression [30] and angiogenesis [28] in the hippocampus of rats. Given that alterations of brain capillaries have been observed in schizophrenia [154], the demonstrated angiogenic effect of risperidone might be a partial mechanism by which antipsychotics realized their action.

Deficit of parvalbumin-expressing interneurons is an accepted feature in schizophrenia [148, 155], also demonstrable in animal models [156160]. Our findings that risperidone counteracts this induced by prenatal poly I:C deficiency may also have important implications for understanding its antipsychotic mechanism.

To fully complete the present section, it is important to note that, in line with previously reported findings [161, 162], we found no effects of prenatal poly I:C treatment and/or risperidone intervention on astrocytes, oligodendrocytes, and microglial cells.

10. Third Lesson: The Aspect of Imbalances in Central Glutamatergic Neurotransmission

Glutamatergic alterations have been consistently showed in psychosis, since its earliest stages [54]. Repeated administration of noncompetitive NMDA receptor antagonists like ketamine, dizocilpine, and phencyclidine (PCP) to neonatal and pubertal rats leads to a number of molecular, neurochemical, and behavioural alterations that resemble those observed in schizophrenia [156, 163, 164]. Administration of the NMDA receptor antagonist dizocilpine and PCP in late fetal and early postnatal period of life in the rat will increase neuronal death by apoptosis [165]. On the contrary, administration of these substances to rats at an adult age will increase neuronal damage by necrosis with subsequent gliosis [166] which results in enduring alteration in the neuronal circuitry. Maeda and coworkers [66] showed that PCP-induced decreased adult neurogenesis was counteracted by coadministered glycine and D-serine confirming the involvement of NMDA receptors in disruption of neurogenesis. Moreover, they were able to demonstrate a reconstruction of neurogenesis by clozapine, but not haloperidol.

However, we found that acute application of ketamine in sub-anaesthetic doses had no effect on cell proliferation. Animals, decapitated 3 weeks after ketamine application, however, showed a significant increased number of BrdU-labelled nuclei in the subgranular zone compared to saline-treated and untreated animals, whereby the cell-type assignment did not differ between the groups. There was no difference between the left and the right hippocampus, but significantly more BrdU-labelled cells were found in the lateral than in the medial blade of the dentate gyrus. In our first respective paper [167], this was interpreted as stimulating effect of ketamine on neurogenesis. Later on [30], however, we speculated that the withdrawal rather than the application of ketamine was essential and that beside an increase of cell proliferation there was a better survival of proliferated cells. These effects were accompanied by an enhanced mRNA level of BDNF.

Haloperidol and the atypical antipsychotic risperidone increased the total number of BrdU-labelled cells surviving for three weeks within the granule cell layer in untreated animals. Hereby, VEGF (vascular endothelial growth factor, signalling protein involved in angiogenesis and cell proliferation in general), MMP2 (matrix metallopeptidase 2 (Gelatinase A)), a proteolytic enzyme involved in cell proliferation, adhesion, and migration), CREB, and p38 MAP kinase seemed to be involved at mRNA as well as protein levels. The ketamine withdrawal-induced changes in proliferation/survival, however, were not additionally affected by the neuroleptics [30].

Malberg and Monteggia [168] showed that chronic administration of haloperidol increased the level of BDNF in the frontal cortex and amygdala, a possible mechanism for the neuroproliferative potency of haloperidol. Together with its direct effect on MMP2 and the subsequent effect on VEGF (possibly by processing the VEGF binding proteins HARP (heparin affine regulatory peptide)) and CTGF (connective tissue growth factor), the cell proliferative/protective potency of haloperidol is plausible.

Nevertheless, the demonstrated haloperidol effect on cell proliferation is in agreement with some, but not all, previous reports [24, 25, 27, 169]. The differences can reflect methodological heterogeneity across different experimental settings (dosage, application regime, and used rat strain).

Interestingly, in the PCP model risperidone was unable to reverse the PCP-induced decreases in parvalbumin expression in the prefrontal cortex [65]. This indicates that the antipsychotic effects of risperidone differ (prenatal poly I:C insult (see above) versus chronic administration of PCP to adult animals).

11. Fourth Lesson: The Aspect of Mechanical Lesions

Lesion models such as the neonatal ventral hippocampal lesion result in schizophrenia-related alterations in behaviour, neurochemistry, and neuropathology when performed on postnatal day (PD) 7, but not on PD 14 or PD 21 [32]. Interestingly, lesions performed in adolescent rats result in less pronounced and qualitatively different schizophrenia-related alterations [170173]. However, only few data suggesting a link between brain lesion and neurogenesis are available. Lipska et al. [174] and Ashe et al. [175] studied the expression of BDNF mRNA in rats with neonatal lesions of the ventral hippocampus and found consistently a suppressed BDNF level in the dentate gyrus. From that and from the BrdU incorporation studies, they concluded that “a transient disconnection in the CA1 and CA2 area of the hippocampus may have long-lasting consequences for neurogenesis in the dentate gyrus” [176]. Negrete-Díaz et al. [67] showed that nitric oxide (NO) levels in the prefrontal cortex, the occipital cortex, and the cerebellum are higher in the damaged animals and that haloperidol, in part, attenuates these altered NO levels. NO itself is known to be anti-proliferative and it should be allowed to suppose a connection between the enhanced NO level and the reduced BrdU incorporation in animals with ventral hippocampus lesions. It is not clear, however, how haloperidol-induced reduction in NO may lead to a restored cell proliferation. This might be, at least partially, a mechanism by which haloperidol decreased stereotypy in ventral hippocampus damaged rats [177].

12. Conclusions/Outlook

Clinical and experimental researches indicate that neurogenesis is disturbed in schizophrenia, since its earliest phases. Moreover, antipsychotics specifically interact with these alterations, affecting the neurogenesis. By increasing the neurogenesis it may be possible to provide beneficial gains for processes related with learning and memory formation. The regulation of neurogenesis may be a promising novel target for the treatment and the prevention of schizophrenia.

References

  • 1.Bayer TA, Falkai P, Maier W. Genetic and non-genetic vulnerability factors in schizophrenia: the basis of the “two hit hypothesis“. Journal of Psychiatric Research. 1999;33(6):543–548. doi: 10.1016/s0022-3956(99)00039-4. [DOI] [PubMed] [Google Scholar]
  • 2.Maynard TM, Sikich L, Lieberman JA, LaMantia AS. Neural development, cell-cell signaling, and the “two-hit” hypothesis of schizophrenia. Schizophrenia Bulletin. 2001;27(3):457–476. doi: 10.1093/oxfordjournals.schbul.a006887. [DOI] [PubMed] [Google Scholar]
  • 3.Pantelis C, Yücel M, Wood SJ, McGorry PD, Velakoulis D. Early and late neurodevelopmental disturbances in schizophrenia and their functional consequences. Australian and New Zealand Journal of Psychiatry. 2003;37(4):399–406. doi: 10.1046/j.1440-1614.2003.01193.x. [DOI] [PubMed] [Google Scholar]
  • 4.Rapoport JL, Addington AM, Frangou S, Psych MRC. The neurodevelopmental model of schizophrenia: update 2005. Molecular Psychiatry. 2005;10(5):434–449. doi: 10.1038/sj.mp.4001642. [DOI] [PubMed] [Google Scholar]
  • 5.Fusar-Poli P, Borgwardt S, Bechdolf A, et al. The psychosis high risk state: a comprehensive state of the art review. doi: 10.1001/jamapsychiatry.2013.269. Archives of General Psychiatry. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Fusar-Poli P, Bonoldi I, Yung AR. Predicting psychosis: meta-analysis of transition outcomes in individuals at high clinical risk. Archives of General Psychiatry. 2012;69(3):220–229. doi: 10.1001/archgenpsychiatry.2011.1472. [DOI] [PubMed] [Google Scholar]
  • 7.Fusar-Poli P, Bechdolf A, Taylor M, et al. At risk for schizophrenic or affective psychosis? A meta-analysis of ICD/DSM diagnostic outcomes in individuals at high clinical risk. doi: 10.1093/schbul/sbs060. Schizophrenia Bulletin. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Fusar-Poli P, Deste G, Smieskova R, et al. Cognitive functioning in prodromal psychosis: a meta-analysis. Archives of General Psychiatry. 2012;69(6):562–571. doi: 10.1001/archgenpsychiatry.2011.1592. [DOI] [PubMed] [Google Scholar]
  • 9.Fusar-Poli P, Borgwardt S, Crescini A, et al. Neuroanatomy of vulnerability to psychosis: a voxel-based meta-analysis. Neuroscience and Biobehavioral Reviews. 2011;35(5):1175–1185. doi: 10.1016/j.neubiorev.2010.12.005. [DOI] [PubMed] [Google Scholar]
  • 10.Fusar-Poli P, Radua J, McGuire P, Borgwardt S. Neuroanatomical maps of psychosis onset: voxel-wise meta-analysis of antipsychotic-naive VBM studies. doi: 10.1093/schbul/sbr134. Schizophrenia Bulletin. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Smieskova R, Fusar-Poli P, Allen P, et al. Neuroimaging predictors of transition to psychosis—a systematic review and meta-analysis. Neuroscience and Biobehavioral Reviews. 2010;34(8):1207–1222. doi: 10.1016/j.neubiorev.2010.01.016. [DOI] [PubMed] [Google Scholar]
  • 12.Fusar-Poli P, Howes OD, Allen P, et al. Abnormal frontostriatal interactions in people with prodromal signs of psychosis: a multimodal imaging study. Archives of General Psychiatry. 2010;67(7):683–691. doi: 10.1001/archgenpsychiatry.2010.77. [DOI] [PubMed] [Google Scholar]
  • 13.Fusar-Poli P, Howes OD, Allen P, et al. Abnormal prefrontal activation directly related to pre-synaptic striatal dopamine dysfunction in people at clinical high risk for psychosis. Molecular Psychiatry. 2011;16(1):67–75. doi: 10.1038/mp.2009.108. [DOI] [PubMed] [Google Scholar]
  • 14.Fusar-Poli P, Perez J, Broome M, et al. Neurofunctional correlates of vulnerability to psychosis: a systematic review and meta-analysis. Neuroscience and Biobehavioral Reviews. 2007;31(4):465–484. doi: 10.1016/j.neubiorev.2006.11.006. [DOI] [PubMed] [Google Scholar]
  • 15.Crossley NA, Mechelli A, Fusar-Poli P, et al. Superior temporal lobe dysfunction and frontotemporal dysconnectivity in subjects at risk of psychosis and in first-episode psychosis. Human Brain Mapping. 2009;30(12):4129–4137. doi: 10.1002/hbm.20834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Fusar-Poli P, Meyer-Lindenberg A. Striatal presynaptic dopamine in schizophrenia—part II: meta-analysis of [18F]/[11C] DOPA PET studies. doi: 10.1093/schbul/sbr180. Schizophrenia Bulletin. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Fusar-Poli P, Stone J, Broome M, Valli I, Mechelli A, McLean M. Thalamic glutamate levels as a predictor of cortical response during executive functioning in subjects at high risk for psychosis. Archives of General Psychiatry. 2011;68(9):881–890. doi: 10.1001/archgenpsychiatry.2011.46. [DOI] [PubMed] [Google Scholar]
  • 18.Howes OD, Montgomery AJ, Asselin MC, et al. Elevated striatal dopamine function linked to prodromal signs of schizophenia. Archives of General Psychiatry. 2009;66(1):13–20. doi: 10.1001/archgenpsychiatry.2008.514. [DOI] [PubMed] [Google Scholar]
  • 19.McGuire P, Howes OD, Stone J, Fusar-Poli P. Functional neuroimaging in schizophrenia: diagnosis and drug discovery. Trends in Pharmacological Sciences. 2008;29(2):91–98. doi: 10.1016/j.tips.2007.11.005. [DOI] [PubMed] [Google Scholar]
  • 20.Fusar-Poli P, Yung AR. Should attenuated psychosis syndrome be included in the DSM5? The debate. The Lancet. 2012;379(9816):591–592. doi: 10.1016/S0140-6736(11)61507-9. [DOI] [PubMed] [Google Scholar]
  • 21.Arnold I, Watt FM. c-Myc activation in transgenic mouse epidermis results in mobilization of stem cells and differentiation of their progeny. Current Biology. 2001;11(8):558–568. doi: 10.1016/s0960-9822(01)00154-3. [DOI] [PubMed] [Google Scholar]
  • 22.Reif A, Fritzen S, Finger M, et al. Neural stem cell proliferation is decreased in schizophrenia, but not in depression. Molecular Psychiatry. 2006;11(5):514–522. doi: 10.1038/sj.mp.4001791. [DOI] [PubMed] [Google Scholar]
  • 23.Kippin TE, Kapur S, Van Der Kooy D. Dopamine specifically inhibits forebrain neural stem cell proliferation, suggesting a novel effect of antipsychotic drugs. Journal of Neuroscience. 2005;25(24):5815–5823. doi: 10.1523/JNEUROSCI.1120-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Wakade CG, Mahadik SP, Waller JL, Chiu FC. Atypical neuroleptics stimulate neurogenesis in adult rat brain. Journal of Neuroscience Research. 2002;69(1):72–79. doi: 10.1002/jnr.10281. [DOI] [PubMed] [Google Scholar]
  • 25.Wang HD, Dunnavant FD, Jarman T, Deutch AY. Effects of antipsychotic drugs on neurogenesis in the forebrain of the adult rat. Neuropsychopharmacology. 2004;29(7):1230–1238. doi: 10.1038/sj.npp.1300449. [DOI] [PubMed] [Google Scholar]
  • 26.Kodama M, Fujioka T, Duman RS. Chronic olanzapine or fluoxetine administration increases cell proliferation in hippocampus and prefrontal cortex of adult rat. Biological Psychiatry. 2004;56(8):570–580. doi: 10.1016/j.biopsych.2004.07.008. [DOI] [PubMed] [Google Scholar]
  • 27.Halim ND, Weickert CS, McClintock BW, Weinberger DR, Lipska BK. Effects of chronic haloperidol and clozapine treatment on neurogenesis in the adult rat hippocampus. Neuropsychopharmacology. 2004;29(6):1063–1069. doi: 10.1038/sj.npp.1300422. [DOI] [PubMed] [Google Scholar]
  • 28.Piontkewitz Y, Bernstein HG, Dobrowolny H, Bogerts B, Weiner I, Keilhoff G. Effects of risperidone treatment in adolescence on hippocampal neurogenesis, parvalbumin expression, and vascularization following prenatal immune activation in rats. Brain, Behavior, and Immunity. 2012;26(2):353–363. doi: 10.1016/j.bbi.2011.11.004. [DOI] [PubMed] [Google Scholar]
  • 29.Keilhoff G, Grecksch G, Becker A. Haloperidol normalized prenatal vitamin D depletion-induced reduction of hippocampal cell proliferation in adult rats. Neuroscience Letters. 2010;476(2):94–98. doi: 10.1016/j.neulet.2010.04.010. [DOI] [PubMed] [Google Scholar]
  • 30.Keilhoff G, Grecksch G, Bernstein HG, Roskoden T, Becker A. Risperidone and haloperidol promote survival of stem cells in the rat hippocampus. European Archives of Psychiatry and Clinical Neuroscience. 2010;260(2):151–162. doi: 10.1007/s00406-009-0033-1. [DOI] [PubMed] [Google Scholar]
  • 31.Castagné V, Moser PC, Porsolt RD. Preclinical behavioral models for predicting antipsychotic activity. Advances in Pharmacology. 2009;57:381–418. doi: 10.1016/S1054-3589(08)57010-4. [DOI] [PubMed] [Google Scholar]
  • 32.Jones CA, Watson DJ, Fone KC. Animal models of schizophrenia. British Journal of Pharmacology. 2011;164(4):1162–1194. doi: 10.1111/j.1476-5381.2011.01386.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Lu L, Mamiya T, Koseki T, Mouri A, Nabeshima T. Genetic animal models of schizophrenia related with the hypothesis of abnormal neurodevelopment. Biological & Pharmaceutical Bulletin. 2011;34(9):1358–1363. doi: 10.1248/bpb.34.1358. [DOI] [PubMed] [Google Scholar]
  • 34.Moore H. The role of rodent models in the discovery of new treatments for schizophrenia: updating our strategy. Schizophrenia Bulletin. 2010;36(6):1066–1072. doi: 10.1093/schbul/sbq106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Neill JC, Barnes S, Cook S, et al. Animal models of cognitive dysfunction and negative symptoms of schizophrenia: focus on NMDA receptor antagonism. Pharmacology and Therapeutics. 2010;128(3):419–432. doi: 10.1016/j.pharmthera.2010.07.004. [DOI] [PubMed] [Google Scholar]
  • 36.Powell SB. Models of neurodevelopmental abnormalities in schizophrenia. Current Topics in Behavioral Neurosciences. 2010;4:435–481. doi: 10.1007/7854_2010_57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Tordjman S, Drapier D, Bonnot O, et al. Animal models relevant to schizophrenia and autism: validity and limitations. Behavior Genetics. 2007;37(1):61–78. doi: 10.1007/s10519-006-9120-5. [DOI] [PubMed] [Google Scholar]
  • 38.Young JW, Zhou X, Geyer MA. Animal models of schizophrenia. Current Topics in Behavioral Neurosciences. 2010;(4):391–433. doi: 10.1007/7854_2010_62. [DOI] [PubMed] [Google Scholar]
  • 39.Ashe PC, Berry MD, Boulton AA. Schizophrenia, a neurodegenerative disorder with neurodevelopmental antecedents. Progress in Neuro-Psychopharmacology and Biological Psychiatry. 2001;25(4):691–707. doi: 10.1016/s0278-5846(01)00159-2. [DOI] [PubMed] [Google Scholar]
  • 40.Cannon M, Murray RM. Neonatal origins of schizophrenia. Archives of Disease in Childhood. 1998;78(1):1–3. doi: 10.1136/adc.78.1.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Weinberger DR. From neuropathology to neurodevelopment. Lancet. 1995;346(8974):552–557. doi: 10.1016/s0140-6736(95)91386-6. [DOI] [PubMed] [Google Scholar]
  • 42.Falkai P, Bogerts B. Cell loss in the hippocampus of schizophrenics. European Archives of Psychiatry and Neurological Sciences. 1986;236(3):154–161. doi: 10.1007/BF00380943. [DOI] [PubMed] [Google Scholar]
  • 43.Heckers S, Konradi C. Hippocampal neurons in schizophrenia. Journal of Neural Transmission. 2002;109(5-6):891–905. doi: 10.1007/s007020200073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Heckers S. The hippocampus in schizophrenia. American Journal of Psychiatry. 2004;161(11):2138–2139. doi: 10.1176/appi.ajp.161.11.2138-a. [DOI] [PubMed] [Google Scholar]
  • 45.Allen NC, Bagade S, McQueen MB, et al. Systematic meta-analyses and field synopsis of genetic association studies in schizophrenia: the SzGene database. Nature Genetics. 2008;40(7):827–834. doi: 10.1038/ng.171. [DOI] [PubMed] [Google Scholar]
  • 46.Desbonnet L, Waddington JL, O’Tuathaigh CMP. Mutant models for genes associated with schizophrenia. Biochemical Society Transactions. 2009;37(1):308–312. doi: 10.1042/BST0370308. [DOI] [PubMed] [Google Scholar]
  • 47.Harrison PJ, Pritchett D, Stumpenhorst K, Betts JF, Nissen W, Schweimer J. Genetic mouse models relevant to schizophrenia: taking stock and looking forward. Neuropharmacology. 2012;62(3):1164–1167. doi: 10.1016/j.neuropharm.2011.08.009. [DOI] [PubMed] [Google Scholar]
  • 48.Guidotti A, Pesold C, Costa E. New neurochemical markers for psychosis: a working hypothesis of their operation. Neurochemical Research. 2000;25(9-10):1207–1218. doi: 10.1023/a:1007635927069. [DOI] [PubMed] [Google Scholar]
  • 49.Stefansson H, Sigurdsson E, Steinthorsdottir V, et al. Neuregulin 1 and susceptibility to schizophrenia. American Journal of Human Genetics. 2002;71(4):877–892. doi: 10.1086/342734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Barros CS, Calabrese B, Chamero P, et al. Impaired maturation of dendritic spines without disorganization of cortical cell layers in mice lacking NRG1/ErbB signaling in the central nervous system. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(11):4507–4512. doi: 10.1073/pnas.0900355106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Niwa M, Kamiya A, Murai R, et al. Knockdown of DISC1 by in utero gene transfer disturbs postnatal dopaminergic maturation in the frontal cortex and leads to adult behavioral deficits. Neuron. 2010;65(4):480–489. doi: 10.1016/j.neuron.2010.01.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Kellendonk C, Simpson EH, Polan HJ, et al. Transient and selective overexpression of dopamine D2 receptors in the striatum causes persistent abnormalities in prefrontal cortex functioning. Neuron. 2006;49(4):603–615. doi: 10.1016/j.neuron.2006.01.023. [DOI] [PubMed] [Google Scholar]
  • 53.Deakin IH, Nissen W, Law AJ, Lane T, Kanso R, Schwab MH. Transgenic overexpression of the type I isoform of neuregulin 1 affects working memory and hippocampal oscillations but not long-term potentiation. Cerebral Cortex. 2012;22(7):1520–1529. doi: 10.1093/cercor/bhr223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Egerton A, Stone JM. The glutamate hypothesis of schizophrenia: neuroimaging and drug development. Current Pharmaceutical Biotechnology. 2012;13(8):1500–1512. doi: 10.2174/138920112800784961. [DOI] [PubMed] [Google Scholar]
  • 55.Duncan GE, Moy SS, Perez A, et al. Deficits in sensorimotor gating and tests of social behavior in a genetic model of reduced NMDA receptor function. Behavioural Brain Research. 2004;153(2):507–519. doi: 10.1016/j.bbr.2004.01.008. [DOI] [PubMed] [Google Scholar]
  • 56.Fradley RL, O’Meara GF, Newman RJ, Andrieux A, Job D, Reynolds DS. STOP knockout and NMDA NR1 hypomorphic mice exhibit deficits in sensorimotor gating. Behavioural Brain Research. 2005;163(2):257–264. doi: 10.1016/j.bbr.2005.05.012. [DOI] [PubMed] [Google Scholar]
  • 57.Halene TB, Ehrlichman RS, Liang Y, et al. Assessment of NMDA receptor NR1 subunit hypofunction in mice as a model for schizophrenia. Genes, Brain and Behavior. 2009;8(7):661–675. doi: 10.1111/j.1601-183X.2009.00504.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Ramsey AJ. NR1 knockdown mice as a representative model of the glutamate hypothesis of schizophrenia. Progress in Brain Research. 2009;179:51–58. doi: 10.1016/S0079-6123(09)17906-2. [DOI] [PubMed] [Google Scholar]
  • 59.Gainetdinov RR. Dopamine transporter mutant mice in experimental neuropharmacology. Naunyn-Schmiedeberg’s Archives of Pharmacology. 2008;377(4–6):301–313. doi: 10.1007/s00210-007-0216-0. [DOI] [PubMed] [Google Scholar]
  • 60.O'Tuathaigh CM, Desbonnet L, Moran PM, Kirby BP, Waddington JL. Molecular genetic models related to schizophrenia and psychotic illness: heuristics and challenges. Current Topics in Behavioral Neurosciences. 2011;7:87–119. doi: 10.1007/7854_2010_111. [DOI] [PubMed] [Google Scholar]
  • 61.O’Tuathaigh CMP, Kirby BP, Moran PM, Waddington JL. Mutant mouse models: genotype-phenotype relationships to negative symptoms in schizophrenia. Schizophrenia Bulletin. 2010;36(2):271–288. doi: 10.1093/schbul/sbp125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Babovic D, O’Tuathaigh CM, O’Connor AM, et al. Phenotypic characterization of cognition and social behavior in mice with heterozygous versus homozygous deletion of catechol-O-methyltransferase. Neuroscience. 2008;155(4):1021–1029. doi: 10.1016/j.neuroscience.2008.07.006. [DOI] [PubMed] [Google Scholar]
  • 63.Gogos JA, Morgan M, Luine V, et al. Catechol-O-methyltransferase-deficient mice exhibit sexually dimorphic changes in catecholamine levels and behavior. Proceedings of the National Academy of Sciences of the United States of America. 1998;95(17):9991–9996. doi: 10.1073/pnas.95.17.9991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Meyer U, Knuesel I, Nyffeler M, Feldon J. Chronic clozapine treatment improves prenatal infection-induced working memory deficits without influencing adult hippocampal neurogenesis. Psychopharmacology. 2010;208(4):531–543. doi: 10.1007/s00213-009-1754-6. [DOI] [PubMed] [Google Scholar]
  • 65.McKibben CE, Jenkins TA, Adams HN, Harte MK, Reynolds GP. Effect of pretreatment with risperidone on phencyclidine-induced disruptions in object recognition memory and prefrontal cortex parvalbumin immunoreactivity in the rat. Behavioural Brain Research. 2010;208(1):132–136. doi: 10.1016/j.bbr.2009.11.018. [DOI] [PubMed] [Google Scholar]
  • 66.Maeda K, Sugino H, Hirose T, et al. Clozapine prevents a decrease in neurogenesis in mice repeatedly treated with phencyclidine. Journal of Pharmacological Sciences. 2007;103(3):299–308. doi: 10.1254/jphs.fp0061424. [DOI] [PubMed] [Google Scholar]
  • 67.Negrete-Díaz J, Baltazar-Gaytán E, Bringas M, et al. Neonatal ventral hippocampus lesion induces increase in no levels which is attenuated by subchronic haloperidol treatment. Synapse. 2010;64(12):941–947. doi: 10.1002/syn.20835. [DOI] [PubMed] [Google Scholar]
  • 68.Gage FH. Mammalian neural stem cells. Science. 2000;287(5457):1433–1438. doi: 10.1126/science.287.5457.1433. [DOI] [PubMed] [Google Scholar]
  • 69.Alvarez-Buylla A, García-Verdugo JM. Neurogenesis in adult subventricular zone. Journal of Neuroscience. 2002;22(3):629–634. doi: 10.1523/JNEUROSCI.22-03-00629.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Ma DK, Marchetto MC, Guo JU, Ming GL, Gage FH, Song H. Epigenetic choreographers of neurogenesis in the adult mammalian brain. Nature Neuroscience. 2010;13(11):1338–1344. doi: 10.1038/nn.2672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Artegiani B, Calegari F. Age-related cognitive decline: can neural stem cells help us? Aging. 2012;4(3):176–186. doi: 10.18632/aging.100446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Marin-Burgin A, Schinder AF. Requirement of adult-born neurons for hippocampus-dependent learning. Behavioural Brain Research. 2012;227(2):391–399. doi: 10.1016/j.bbr.2011.07.001. [DOI] [PubMed] [Google Scholar]
  • 73.Snyder JS, Soumier A, Brewer M, Pickel J, Cameron HA. Adult hippocampal neurogenesis buffers stress responses and depressive behaviour. Nature. 2011;476(7361):458–461. doi: 10.1038/nature10287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Eisch AJ, Cameron HA, Encinas JM, Meltzer LA, Ming GL, Overstreet-Wadiche LS. Adult neurogenesis, mental health, and mental illness: hope or hype? Journal of Neuroscience. 2008;28(46):11785–11791. doi: 10.1523/JNEUROSCI.3798-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Foret A, Quertainmont R, Botman O, et al. Stem cells in the adult rat spinal cord: plasticity after injury and treadmill training exercise. Journal of Neurochemistry. 2010;112(3):762–772. doi: 10.1111/j.1471-4159.2009.06500.x. [DOI] [PubMed] [Google Scholar]
  • 76.Xiong Y, Mahmood A, Chopp M. Angiogenesis, neurogenesis and brain recovery of function following injury. Current Opinion in Investigational Drugs. 2010;11(3):298–308. [PMC free article] [PubMed] [Google Scholar]
  • 77.Erlandsson A, Lin CH, Yu F, Morshead CM. Immunosuppression promotes endogenous neural stem and progenitor cell migration and tissue regeneration after ischemic injury. Experimental Neurology. 2011;230(1):48–57. doi: 10.1016/j.expneurol.2010.05.018. [DOI] [PubMed] [Google Scholar]
  • 78.Zhao C, Zawadzka M, Roulois AJA, Bruce CC, Franklin RJM. Promoting remyelination in multiple sclerosis by endogenous adult neural stem/precursor cells: defining cellular targets. Journal of the Neurological Sciences. 2008;265(1-2):12–16. doi: 10.1016/j.jns.2007.05.008. [DOI] [PubMed] [Google Scholar]
  • 79.Hugnot JP, Franzen R. The spinal cord ependymal region: a stem cell niche in the caudal central nervous system. Frontiers in Bioscience. 2011;16(3):1044–1059. doi: 10.2741/3734. [DOI] [PubMed] [Google Scholar]
  • 80.Geraerts M, Krylyshkina O, Debyser Z, Baekelandt V. Concise review: therapeutic strategies for Parkinson disease based on the modulation of adult neurogenesis. Stem Cells. 2007;25(2):263–270. doi: 10.1634/stemcells.2006-0364. [DOI] [PubMed] [Google Scholar]
  • 81.Rioux L, Arnold SE. The expression of retinoic acid receptor alpha is increased in the granule cells of the dentate gyrus in schizophrenia. Psychiatry Research. 2005;133(1):13–21. doi: 10.1016/j.psychres.2004.11.003. [DOI] [PubMed] [Google Scholar]
  • 82.Atz ME, Rollins B, Vawter MP. NCAM1 association study of bipolar disorder and schizophrenia: polymorphisms and alternatively spliced isoforms lead to similarities and differences. Psychiatric Genetics. 2007;17(2):55–67. doi: 10.1097/YPG.0b013e328012d850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Gascon E, Vutskits L, Kiss JZ. The role of PSA-NCAM in adult neurogenesis. Advances in Experimental Medicine and Biology. 2010;663:127–136. doi: 10.1007/978-1-4419-1170-4_8. [DOI] [PubMed] [Google Scholar]
  • 84.Bernstein HG, Steiner J, Bogerts B. Glial cells in schizophrenia: pathophysiological significance and possible consequences for therapy. Expert Review of Neurotherapeutics. 2009;9(7):1059–1071. doi: 10.1586/ern.09.59. [DOI] [PubMed] [Google Scholar]
  • 85.Fields RD. White matter in learning, cognition and psychiatric disorders. Trends in Neurosciences. 2008;31(7):361–370. doi: 10.1016/j.tins.2008.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Takahashi N, Sakurai T, Davis KL, Buxbaum JD. Linking oligodendrocyte and myelin dysfunction to neurocircuitry abnormalities in schizophrenia. Progress in Neurobiology. 2010;93(1):13–22. doi: 10.1016/j.pneurobio.2010.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Rothermundt M, Ahn JN, Jorgens S. S100B in schizophrenia: an update. General Physiology and Biophysics. 2009;28:F76–F81. [PubMed] [Google Scholar]
  • 88.Bilbo SD, Schwarz JM. Early-life programming of later-life brain and behavior: a critical role for the immune system. Frontiers in Behavioral Neuroscience. 2009;3, article 14 doi: 10.3389/neuro.08.014.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Harry GJ, Kraft AD. Microglia in the developing brain: a potential target with lifetime effects. Neurotoxicology. 2012;33(2):191–206. doi: 10.1016/j.neuro.2012.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Neumann H, Kotter MR, Franklin RJM. Debris clearance by microglia: an essential link between degeneration and regeneration. Brain. 2009;132(2):288–295. doi: 10.1093/brain/awn109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Olah M, Amor S, Brouwer N, Vinet J, Eggen B, Biber K. Identification of a microglia phenotype supportive of remyelination. Glia. 2012;60(2):306–321. doi: 10.1002/glia.21266. [DOI] [PubMed] [Google Scholar]
  • 92.Tremblay ME, Stevens B, Sierra A, Wake H, Bessis A. Nimmerjahn A. The role of microglia in the healthy brain. Journal of Neuroscience. 2011;31(45):16064–16069. doi: 10.1523/JNEUROSCI.4158-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Crossley NA, Constante M, McGuire P, Power P. Efficacy of atypical v. typical antipsychotics in the treatment of early psychosis: meta-analysis. British Journal of Psychiatry. 2010;196(6):434–439. doi: 10.1192/bjp.bp.109.066217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Horacek J, Bubenikova-Valesova V, Kopecek M, et al. Mechanism of action of atypical antipsychotic drugs and the neurobiology of schizophrenia. CNS Drugs. 2006;20(5):389–409. doi: 10.2165/00023210-200620050-00004. [DOI] [PubMed] [Google Scholar]
  • 95.Angelucci F, Mathé AA, Aloe L. Neurotrophic factors and CNS disorders: findings in rodent models of depression and schizophrenia. Progress in Brain Research. 2004;146:151–165. doi: 10.1016/s0079-6123(03)46011-1. [DOI] [PubMed] [Google Scholar]
  • 96.Prommer E. Olanzapine: palliative medicine update. doi: 10.1177/1049909112441241. American Journal of Hospice Palliative Care. In press. [DOI] [PubMed] [Google Scholar]
  • 97.Steen RG, Mull C, McClure R, Hamer RM, Lieberman JA. Brain volume in first-episode schizophrenia: systematic review and meta-analysis of magnetic resonance imaging studies. British Journal of Psychiatry. 2006;188:510–518. doi: 10.1192/bjp.188.6.510. [DOI] [PubMed] [Google Scholar]
  • 98.Selemon LD, Lidow MS, Goldman-Rakic PS. Increased volume and glial density in primate prefrontal cortex associated with chronic antipsychotic drug exposure. Biological Psychiatry. 1999;46(2):161–172. doi: 10.1016/s0006-3223(99)00113-4. [DOI] [PubMed] [Google Scholar]
  • 99.Riva MA, Molteni R, Tascedda F, Massironi A, Racagni G. Selective modulation of fibroblast growth factor-2 expression in the rat brain by the atypical antipsychotic clozapine. Neuropharmacology. 1999;38(7):1075–1082. doi: 10.1016/s0028-3908(99)00031-3. [DOI] [PubMed] [Google Scholar]
  • 100.Fumagalli F, Molteni R, Bedogni F, et al. Quetiapine regulates FGF-2 and BDNF expression in the hippocampus of animals treated with MK-801. NeuroReport. 2004;15(13):2109–2112. doi: 10.1097/00001756-200409150-00022. [DOI] [PubMed] [Google Scholar]
  • 101.Angelucci F, Brenè S, Mathé AA. BDNF in schizophrenia, depression and corresponding animal models. Molecular Psychiatry. 2005;10(4):345–352. doi: 10.1038/sj.mp.4001637. [DOI] [PubMed] [Google Scholar]
  • 102.Buckley PF, Pillai A, Evans D, Stirewalt E, Mahadik S. Brain derived neurotropic factor in first-episode psychosis. Schizophrenia Research. 2007;91(1–3):1–5. doi: 10.1016/j.schres.2006.12.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Buckley PF, Pillai A, Howell KR. Brain-derived neurotrophic factor: findings in schizophrenia. Current Opinion in Psychiatry. 2011;24(2):122–127. doi: 10.1097/YCO.0b013e3283436eb7. [DOI] [PubMed] [Google Scholar]
  • 104.Green MJ, Matheson SL, Shepherd A, Weickert CS, Carr VJ. Brain-derived neurotrophic factor levels in schizophrenia: a systematic review with meta-analysis. Molecular Psychiatry. 2011;16(9):960–972. doi: 10.1038/mp.2010.88. [DOI] [PubMed] [Google Scholar]
  • 105.Guillin O, Demily C, Thibaut F. Brain-derived neurotrophic factor in schizophrenia and its relation with dopamine. International Review of Neurobiology. 2007;78:377–395. doi: 10.1016/S0074-7742(06)78012-6. [DOI] [PubMed] [Google Scholar]
  • 106.Chlan-Fourney J, Ashe P, Nylen K, Juorio AV, Li XM. Differential regulation of hippocampal BDNF mRNA by typical and atypical antipsychotic administration. Brain Research. 2002;954(1):11–20. doi: 10.1016/s0006-8993(02)03215-8. [DOI] [PubMed] [Google Scholar]
  • 107.Dawson NM, Hamid EH, Egan MF, Meredith GE. Changes in the pattern of brain-derived neurotrophic factor immunoreactivity in the rat brain after acute and subchronic haloperidol treatment. Synapse. 2001;39(1):70–81. doi: 10.1002/1098-2396(20010101)39:1<70::AID-SYN10>3.0.CO;2-J. [DOI] [PubMed] [Google Scholar]
  • 108.Lindén AM, Väisänen J, Lakso M, Nawa H, Wong G, Castrén E. Expression of neurotrophins BDNF and NT-3, and their receptors in rat brain after administration of antipsychotic and psychotrophic agents. Journal of Molecular Neuroscience. 2000;14(1-2):27–37. doi: 10.1385/JMN:14:1-2:027. [DOI] [PubMed] [Google Scholar]
  • 109.Parikh V, Khan MM, Mahadik SP. Olanzapine counteracts reduction of brain-derived neurotrophic factor and TrkB receptors in rat hippocampus produced by haloperidol. Neuroscience Letters. 2004;356(2):135–139. doi: 10.1016/j.neulet.2003.10.079. [DOI] [PubMed] [Google Scholar]
  • 110.Parikh V, Terry AV, Khan MM, Mahadik SP. Modulation of nerve growth factor and choline acetyltransferase expression in rat hippocampus after chronic exposure to haloperidol, risperidone, and olanzapine. Psychopharmacology. 2004;172(4):365–374. doi: 10.1007/s00213-003-1669-6. [DOI] [PubMed] [Google Scholar]
  • 111.Terry AV, Jr., Gearhart DA, Warner S, et al. Protracted effects of chronic oral haloperidol and risperidone on nerve growth factor, cholinergic neurons, and spatial reference learning in rats. Neuroscience. 2007;150(2):413–424. doi: 10.1016/j.neuroscience.2007.09.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Pillai A, Mahadik SP. Differential effects of haloperidol and olanzapine on levels of vascular endothelial growth factor and angiogenesis in rat hippocampus. Schizophrenia Research. 2006;87(1–3):48–59. doi: 10.1016/j.schres.2006.06.017. [DOI] [PubMed] [Google Scholar]
  • 113.Pozzi L, Håkansson K, Usiello A, et al. Opposite regulation by typical and atypical anti-psychotics of ERK1/2, CREB and Elk-1 phosphorylation in mouse dorsal striatum. Journal of Neurochemistry. 2003;86(2):451–459. doi: 10.1046/j.1471-4159.2003.01851.x. [DOI] [PubMed] [Google Scholar]
  • 114.Yang BH, Son H, Hyeon Kim S, Nam JH, Choi JH, Lee JS. Phosphorylation of ERK and CREB in cultured hippocampal neurons after haloperidol and risperidone administration. Psychiatry and Clinical Neurosciences. 2004;58(3):262–267. doi: 10.1111/j.1440-1819.2004.01229.x. [DOI] [PubMed] [Google Scholar]
  • 115.Freyberg Z, Ferrando SJ, Javitch JA. Roles of the Akt/GSK-3 and Wnt signaling pathways in schizophrenia and antipsychotic drug action. American Journal of Psychiatry. 2010;167(4):388–396. doi: 10.1176/appi.ajp.2009.08121873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Crowley JJ, Kim Y, Szatkiewicz JP, Pratt AL, Quackenbush CR, Adkins DE. Genome-wide association mapping of loci for antipsychotic-induced extrapyramidal symptoms in mice. Mammalian Genome. 2012;23(5-6):322–335. doi: 10.1007/s00335-011-9385-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Putnam DK, Sun J, Zhao Z. Exploring schizophrenia drug-gene interactions through molecular network and pathway modeling. Proceedings of the AMIA Annual Symposium; 2011; pp. 1127–1133. [PMC free article] [PubMed] [Google Scholar]
  • 118.Nacher J, McEwen BS. The role of N-methyl-D-asparate receptors in neurogenesis. Hippocampus. 2006;16(3):267–270. doi: 10.1002/hipo.20160. [DOI] [PubMed] [Google Scholar]
  • 119.Archer T, Beninger RJ, Palomo T, Kostrzewa RM. Epigenetics and biomarkers in the staging of neuropsychiatric disorders. Neurotoxicity Research. 2010;18(3-4):347–366. doi: 10.1007/s12640-010-9163-5. [DOI] [PubMed] [Google Scholar]
  • 120.Maric NP, Svrakic DM. Why schizophrenia genetics needs epigenetics: a review. Psychiatria Danubina. 2012;24(1):2–18. [PubMed] [Google Scholar]
  • 121.Sun J, Sun J, Ming GL, Song H. Epigenetic regulation of neurogenesis in the adult mammalian brain. European Journal of Neuroscience. 2011;33(6):1087–1093. doi: 10.1111/j.1460-9568.2011.07607.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Newton SS, Duman RS. Neurogenic actions of atypical antipsychotic drugs and therapeutic implications. CNS Drugs. 2007;21(9):715–725. doi: 10.2165/00023210-200721090-00002. [DOI] [PubMed] [Google Scholar]
  • 123.Meyer U, Feldon J. Epidemiology-driven neurodevelopmental animal models of schizophrenia. Progress in Neurobiology. 2010;90(3):285–326. doi: 10.1016/j.pneurobio.2009.10.018. [DOI] [PubMed] [Google Scholar]
  • 124.Trivedi MS, Jarbe T. A brief review on recent developments in animal models of schizophrenia. Indian Journal of Pharmacology. 2011;43(4):375–380. doi: 10.4103/0253-7613.83104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Yanagi M, Southcott S, Lister J, Tamminga CA. Animal models of schizophrenia emphasizing construct validity. Progress in Molecular Biology and Translational Science. 2012;105:411–444. doi: 10.1016/B978-0-12-394596-9.00012-3. [DOI] [PubMed] [Google Scholar]
  • 126.Wilson C, Terry AV., Jr. Neurodevelopmental animal models of schizophrenia: role in novel drug discovery and development. Clinical Schizophrenia and Related Psychoses. 2010;4(2):124–137. doi: 10.3371/CSRP.4.2.4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Kesby JP, Eyles DW, Burne THJ, McGrath JJ. The effects of vitamin D on brain development and adult brain function. Molecular and Cellular Endocrinology. 2011;347(1-2):121–127. doi: 10.1016/j.mce.2011.05.014. [DOI] [PubMed] [Google Scholar]
  • 128.McGrath JJ, Burne TH, Féron F, MacKay-Sim A, Eyles DW. Developmental vitamin D deficiency and risk of schizophrenia: a 10-year update. Schizophrenia Bulletin. 2010;36(6):1073–1078. doi: 10.1093/schbul/sbq101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.O’Loan J, Eyles DW, Kesby J, Ko P, McGrath JJ, Burne THJ. Vitamin D deficiency during various stages of pregnancy in the rat; its impact on development and behaviour in adult offspring. Psychoneuroendocrinology. 2007;32(3):227–234. doi: 10.1016/j.psyneuen.2006.12.006. [DOI] [PubMed] [Google Scholar]
  • 130.Eyles D, Brown J, Mackay-Sim A, McGrath J, Feron F. Vitamin D3 and brain development. Neuroscience. 2003;118(3):641–653. doi: 10.1016/s0306-4522(03)00040-x. [DOI] [PubMed] [Google Scholar]
  • 131.Ko P, Burkert R, McGrath J, Eyles D. Maternal vitamin D3 deprivation and the regulation of apoptosis and cell cycle during rat brain development. Developmental Brain Research. 2004;153(1):61–68. doi: 10.1016/j.devbrainres.2004.07.013. [DOI] [PubMed] [Google Scholar]
  • 132.Becker A, Eyles DW, McGrath JJ, Grecksch G. Transient prenatal vitamin D deficiency is associated with subtle alterations in learning and memory functions in adult rats. Behavioural Brain Research. 2005;161(2):306–312. doi: 10.1016/j.bbr.2005.02.015. [DOI] [PubMed] [Google Scholar]
  • 133.Kempermann G, Gage FH. Neurogenesis in the adult hippocampus. Proceedings of the Novartis Foundation Symposium; 2000; pp. 220–235. [PubMed] [Google Scholar]
  • 134.Meltzer HY, Park S, Kessler R. Cognition, schizophrenia, and the atypical antipsychotic drugs. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(24):13591–13593. doi: 10.1073/pnas.96.24.13591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Mishara AL, Goldberg TE. A meta-analysis and critical review of the effects of conventional neuroleptic treatment on cognition in schizophrenia: opening a closed book. Biological Psychiatry. 2004;55(10):1013–1022. doi: 10.1016/j.biopsych.2004.01.027. [DOI] [PubMed] [Google Scholar]
  • 136.Becker A, Grecksch G. Pharmacological treatment to augment hole board habituation in prenatal Vitamin D-deficient rats. Behavioural Brain Research. 2006;166(1):177–183. doi: 10.1016/j.bbr.2005.07.010. [DOI] [PubMed] [Google Scholar]
  • 137.Féron F, Burne THJ, Brown J, et al. Developmental Vitamin D3 deficiency alters the adult rat brain. Brain Research Bulletin. 2005;65(2):141–148. doi: 10.1016/j.brainresbull.2004.12.007. [DOI] [PubMed] [Google Scholar]
  • 138.Keilhoff G. nNOS deficiency-induced cell proliferation depletes the neurogenic reserve. Neuroscience Letters. 2011;505(3):248–253. doi: 10.1016/j.neulet.2011.10.026. [DOI] [PubMed] [Google Scholar]
  • 139.Fatemi SH, Emamian ES, Kist D, et al. Defective corticogenesis and reduction in Reelin immunoreactivity in cortex and hippocampus of prenatally infected neonatal mice. Molecular Psychiatry. 1999;4(2):145–154. doi: 10.1038/sj.mp.4000520. [DOI] [PubMed] [Google Scholar]
  • 140.Macedo DS, Araujo DP, Sampaio LR, Vasconcelos SM, Sales PM, Sousa FC. Animal models of prenatal immune challenge and their contribution to the study of schizophrenia: a systematic review. Brazilian Journal of Medical and Biological Research. 2012;45(3):179–186. doi: 10.1590/S0100-879X2012007500031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Cui K, Ashdown H, Luheshi GN, Boksa P. Effects of prenatal immune activation on hippocampal neurogenesis in the rat. Schizophrenia Research. 2009;113(2-3):288–297. doi: 10.1016/j.schres.2009.05.003. [DOI] [PubMed] [Google Scholar]
  • 142.Lanté F, Meunier J, Guiramand J, et al. Late N-acetylcysteine treatment prevents the deficits induced in the offspring of dams exposed to an immune stress during gestation. Hippocampus. 2008;18(6):602–609. doi: 10.1002/hipo.20421. [DOI] [PubMed] [Google Scholar]
  • 143.Meyer U, Nyffeler M, Yee BK, Knuesel I, Feldon J. Adult brain and behavioral pathological markers of prenatal immune challenge during early/middle and late fetal development in mice. Brain, Behavior, and Immunity. 2008;22(4):469–486. doi: 10.1016/j.bbi.2007.09.012. [DOI] [PubMed] [Google Scholar]
  • 144.Wolf SA, Melnik A, Kempermann G. Physical exercise increases adult neurogenesis and telomerase activity, and improves behavioral deficits in a mouse model of schizophrenia. Brain, Behavior, and Immunity. 2011;25(5):971–980. doi: 10.1016/j.bbi.2010.10.014. [DOI] [PubMed] [Google Scholar]
  • 145.Kempermann G, Krebs J, Fabel K. The contribution of failing adult hippocampal neurogenesis to psychiatric disorders. Current Opinion in Psychiatry. 2008;21(3):290–295. doi: 10.1097/YCO.0b013e3282fad375. [DOI] [PubMed] [Google Scholar]
  • 146.Reif A, Schmitt A, Fritzen S, Lesch KP. Neurogenesis and schizophrenia: dividing neurons in a divided mind? European Archives of Psychiatry and Clinical Neuroscience. 2007;257(5):290–299. doi: 10.1007/s00406-007-0733-3. [DOI] [PubMed] [Google Scholar]
  • 147.Toro CT, Deakin JFW. Adult neurogenesis and schizophrenia: a window on abnormal early brain development? Schizophrenia Research. 2007;90(1–3):1–14. doi: 10.1016/j.schres.2006.09.030. [DOI] [PubMed] [Google Scholar]
  • 148.Zhang ZJ, Reynolds GP. A selective decrease in the relative density of parvalbumin-immunoreactive neurons in the hippocampus in schizophrenia. Schizophrenia Research. 2002;55(1-2):1–10. doi: 10.1016/s0920-9964(01)00188-8. [DOI] [PubMed] [Google Scholar]
  • 149.Piontkewitz Y, Arad M, Weiner I. Risperidone administered during asymptomatic period of adolescence prevents the emergence of brain structural pathology and behavioral abnormalities in an animal model of schizophrenia. Schizophrenia Bulletin. 2011;37(6):1257–1269. doi: 10.1093/schbul/sbq040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Piontkewitz Y, Assaf Y, Weiner I. Clozapine administration in adolescence prevents postpubertal emergence of brain structural pathology in an animal model of schizophrenia. Biological Psychiatry. 2009;66(11):1038–1046. doi: 10.1016/j.biopsych.2009.07.005. [DOI] [PubMed] [Google Scholar]
  • 151.Fournier NM, Duman RS. Role of vascular endothelial growth factor in adult hippocampal neurogenesis: implications for the pathophysiology and treatment of depression. Behavioural Brain Research. 2012;227(2):p. 440, p. 449. doi: 10.1016/j.bbr.2011.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Louissaint A, Jr., Rao S, Leventhal C, Goldman SA. Coordinated interaction of neurogenesis and angiogenesis in the adult songbird brain. Neuron. 2002;34(6):945–960. doi: 10.1016/s0896-6273(02)00722-5. [DOI] [PubMed] [Google Scholar]
  • 153.Schänzer A, Wachs FP, Wilhelm D, et al. Direct stimulation of adult neural stem cells in vitro and neurogenesis in vivo by vascular endothelial growth factor. Brain Pathology. 2004;14(3):237–248. doi: 10.1111/j.1750-3639.2004.tb00060.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Uranova NA, Zimina IS, Vikhreva OV, Krukov NO, Rachmanova VI, Orlovskaya DD. Ultrastructural damage of capillaries in the neocortex in schizophrenia. The World Journal of Biological Psychiatry. 2010;11(3):567–578. doi: 10.3109/15622970903414188. [DOI] [PubMed] [Google Scholar]
  • 155.Lewis DA, Hashimoto T, Volk DW. Cortical inhibitory neurons and schizophrenia. Nature Reviews Neuroscience. 2005;6(4):312–324. doi: 10.1038/nrn1648. [DOI] [PubMed] [Google Scholar]
  • 156.Abdul-Monim Z, Neill JC, Reynolds GP. Sub-chronic psychotomimetic phencyclidine induces deficits in reversal learning and alterations in parvalbumin-immunoreactive expression in the rat. Journal of Psychopharmacology. 2007;21(2):198–205. doi: 10.1177/0269881107067097. [DOI] [PubMed] [Google Scholar]
  • 157.Berretta S, Lange N, Bhattacharyya S, Sebro R, Garces J, Benes FM. Long-term effects of amygdala GABA receptor blockade on specific subpopulations of hippocampal interneurons. Hippocampus. 2004;14(7):876–894. doi: 10.1002/hipo.20002. [DOI] [PubMed] [Google Scholar]
  • 158.Harte MK, Powell SB, Swerdlow NR, Geyer MA, Reynolds GP. Deficits in parvalbumin and calbindin immunoreactive cells in the hippocampus of isolation reared rats. Journal of Neural Transmission. 2007;114(7):893–898. doi: 10.1007/s00702-007-0627-6. [DOI] [PubMed] [Google Scholar]
  • 159.Lodge DJ, Behrens MM, Grace AA. A loss of parvalbumin-containing interneurons is associated with diminished oscillatory activity in an animal model of schizophrenia. Journal of Neuroscience. 2009;29(8):2344–2354. doi: 10.1523/JNEUROSCI.5419-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Penschuck S, Flagstad P, Didriksen M, Leist M, Michael-Titus AT. Decrease in parvalbumin-expressing neurons in the hippocampus and increased phencyclidine-induced locomotor activity in the rat methylazoxymethanol (MAM) model of schizophrenia. European Journal of Neuroscience. 2006;23(1):279–284. doi: 10.1111/j.1460-9568.2005.04536.x. [DOI] [PubMed] [Google Scholar]
  • 161.Makinodan M, Tatsumi K, Manabe T, et al. Maternal immune activation in mice delays myelination and axonal development in the hippocampus of the offspring. Journal of Neuroscience Research. 2008;86(10):2190–2200. doi: 10.1002/jnr.21673. [DOI] [PubMed] [Google Scholar]
  • 162.Nyffeler M, Meyer U, Yee BK, Feldon J, Knuesel I. Maternal immune activation during pregnancy increases limbic GABAA receptor immunoreactivity in the adult offspring: implications for schizophrenia. Neuroscience. 2006;143(1):51–62. doi: 10.1016/j.neuroscience.2006.07.029. [DOI] [PubMed] [Google Scholar]
  • 163.Becker A, Grecksch G, Schwegler H, Roskoden T. Expression of mRNA of neurotrophic factors and their receptors are significantly altered after subchronic ketamine treatment. Medicinal Chemistry. 2008;4(3):256–263. doi: 10.2174/157340608784325124. [DOI] [PubMed] [Google Scholar]
  • 164.Becker A, Peters B, Schroeder H, Mann T, Huether G, Grecksch G. Ketamine-induced changes in rat behaviour: a possible animal model of schizophrenia. Progress in Neuro-Psychopharmacology and Biological Psychiatry. 2003;27(4):687–700. doi: 10.1016/S0278-5846(03)00080-0. [DOI] [PubMed] [Google Scholar]
  • 165.Ikonomidou C, Bosch F, Miksa M, et al. Blockade of NMDA receptors and apoptotic neurodegeneration in the developing brain. Science. 1999;283(5398):70–74. doi: 10.1126/science.283.5398.70. [DOI] [PubMed] [Google Scholar]
  • 166.Olney JW, Labruyere J, Wang G, Wozniak DF, Price MT, Sesma MA. NMDA antagonist neurotoxicity: mechanism and prevention. Science. 1991;254(5037):1515–1518. doi: 10.1126/science.1835799. [DOI] [PubMed] [Google Scholar]
  • 167.Keilhoff G, Bernstein HG, Becker A, Grecksch G, Wolf G. Increased neurogenesis in a rat ketamine model of schizophrenia. Biological Psychiatry. 2004;56(5):317–322. doi: 10.1016/j.biopsych.2004.06.010. [DOI] [PubMed] [Google Scholar]
  • 168.Malberg JE, Monteggia LM. VGF, a new player in antidepressant action? Science Signaling. 2008;1(18, article pe19) doi: 10.1126/stke.118pe19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Malberg JE, Eisch AJ, Nestler EJ, Duman RS. Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. Journal of Neuroscience. 2000;20(24):9104–9110. doi: 10.1523/JNEUROSCI.20-24-09104.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Becker A, Grecksch G, Bernstein HG, Höllt V, Bogerts B. Social behaviour in rats lesioned with ibotenic acid in the hippocampus: quantitative and qualitative analysis. Psychopharmacology. 1999;144(4):333–338. doi: 10.1007/s002130051015. [DOI] [PubMed] [Google Scholar]
  • 171.Daenen EWPM, Van der Heyden JA, Kruse CG, Wolterink G, Van Ree JM. Adaptation and habituation to an open field and responses to various stressful events in animals with neonatal lesions in the amygdala or ventral hippocampus. Brain Research. 2001;918(1-2):153–165. doi: 10.1016/s0006-8993(01)02987-0. [DOI] [PubMed] [Google Scholar]
  • 172.Tseng KY, Chambers RA, Lipska BK. The neonatal ventral hippocampal lesion as a heuristic neurodevelopmental model of schizophrenia. Behavioural Brain Research. 2009;204(2):295–305. doi: 10.1016/j.bbr.2008.11.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Wood GK, Lipska BK, Weinberger DR. Behavioral changes in rats with early ventral hippocampal damage vary with age at damage. Developmental Brain Research. 1997;101(1-2):17–25. doi: 10.1016/s0165-3806(97)00050-3. [DOI] [PubMed] [Google Scholar]
  • 174.Lipska BK, Khaing ZZ, Weickert CS, Weinberger DR. BDNF mRNA expression in rat hippocampus and prefrontal cortex: effects of neonatal ventral hippocampal damage and antipsychotic drugs. European Journal of Neuroscience. 2001;14(1):135–144. doi: 10.1046/j.1460-9568.2001.01633.x. [DOI] [PubMed] [Google Scholar]
  • 175.Ashe PC, Chlan-Fourney J, Juorio AV, Li XM. Brain-derived neurotrophic factor (BDNF) mRNA in rats with neonatal ibotenic acid lesions of the ventral hippocampus. Brain Research. 2002;956(1):126–135. doi: 10.1016/s0006-8993(02)03176-1. [DOI] [PubMed] [Google Scholar]
  • 176.Lipska BK. Using animal models to test a neurodevelopmental hypothesis of schizophrenia. Journal of Psychiatry and Neuroscience. 2004;29(4):282–286. [PMC free article] [PubMed] [Google Scholar]
  • 177.Al-Amin HA, Weinberger DR, Lipska BK. Exaggerated MK-801-induced motor hyperactivity in rats with the neonatal lesion of the ventral hippocampus. Behavioural Pharmacology. 2000;11(3-4):269–278. doi: 10.1097/00008877-200006000-00010. [DOI] [PubMed] [Google Scholar]

Articles from Neural Plasticity are provided here courtesy of Wiley

RESOURCES