全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Immature Dentate Gyrus: An Endophenotype of Neuropsychiatric Disorders

DOI: 10.1155/2013/318596

Full-Text   Cite this paper   Add to My Lib

Abstract:

Adequate maturation of neurons and their integration into the hippocampal circuit is crucial for normal cognitive function and emotional behavior, and disruption of this process could cause disturbances in mental health. Previous reports have shown that mice heterozygous for a null mutation in α-CaMKII, which encodes a key synaptic plasticity molecule, display abnormal behaviors related to schizophrenia and other psychiatric disorders. In these mutants, almost all neurons in the dentate gyrus are arrested at a pseudoimmature state at the molecular and electrophysiological levels, a phenomenon defined as “immature dentate gyrus (iDG).” To date, the iDG phenotype and shared behavioral abnormalities (including working memory deficit and hyperlocomotor activity) have been discovered in Schnurri-2 knockout, mutant SNAP-25 knock-in, and forebrain-specific calcineurin knockout mice. In addition, both chronic fluoxetine treatment and pilocarpine-induced seizures reverse the neuronal maturation, resulting in the iDG phenotype in wild-type mice. Importantly, an iDG-like phenomenon was observed in post-mortem analysis of brains from patients with schizophrenia/bipolar disorder. Based on these observations, we proposed that the iDG is a potential endophenotype shared by certain types of neuropsychiatric disorders. This review summarizes recent data describing this phenotype and discusses the data’s potential implication in elucidating the pathophysiology of neuropsychiatric disorders. 1. Introduction The exact mechanisms within the brain that underlie most psychiatric disorders remain largely unknown, and one of the major challenges in psychiatric research is to identify the pathophysiology in the brains of patients with these disorders. This is challenging because psychiatric disorders are diagnosed on the basis of behavioral characteristics and not biological criteria. Therefore, each psychiatric disorder likely consists of multiple biologically heterogeneous populations, which further complicates the search for underlying pathophysiologies. Previously, studies have identified the “immature dentate gyrus (iDG),” a potential brain endophenotype shared by several psychiatric disorders, including schizophrenia and bipolar disorder. The iDG was identified in animal models of psychiatric disorders, which were selected using large-scale behavioral screening of genetically engineered mice [1]. In the iDG phenotype, most of the granule cells or principal neurons in the dentate gyrus (DG) within the hippocampus are arrested at a pseudoimmature status, in which the

References

[1]  N. Yamasaki, M. Maekawa, K. Kobayashi et al., “Alpha-CaMKII deficiency causes immature dentate gyrus, a novel candidate endophenotype of psychiatric disorders,” Molecular Brain, vol. 1, article 6, 2008.
[2]  K. Takao, K. Kobayashi, H. Hagihara, et al., “Deficiency of Schnurri-2, an MHC enhancer binding protein, induces mild chronic inflammation in the brain and confers molecular, neuronal, and behavioral phenotypes related to schizophrenia,” Neuropsychopharmacology, 2013.
[3]  K. Ohira, K. Kobayashi, K. Toyama, et al., “Synaptosomal-associated protein 25 mutation converts dentate granule cells to an immature state in adult mice,” Molecular Brain, vol. 6, article 12, 2013.
[4]  H. Hagihara, H. K. Nakamura, K. Toyama, I. A. Graef, G. R. Crabtree, and T. Miyakawa, “Forebrain-specific calcineurin deficiency causes immaturity of the dentate granule cells in adult mice,” SfN meeting 2011 abstract.
[5]  K. Kobayashi, Y. Ikeda, A. Sakai et al., “Reversal of hippocampal neuronal maturation by serotonergic antidepressants,” Proceedings of the National Academy of Sciences of the United States of America, vol. 107, no. 18, pp. 8434–8439, 2010.
[6]  R. Shin, K. Kobayashi, H. Hagihara, et al., “The immature dentate gyrus represents a common endophenotype of psychiatric disorders and epilepsy,” Bipolar Disorders, 2013.
[7]  N. M. Walton, Y. Zhou, J. H. Kogan, et al., “Detection of an immature dentate gyrus feature in human schizophrenia/bipolar patients,” Translational Psychiatry, vol. 2, no. 7, article e135, 2012.
[8]  K. Takao, N. Yamasaki, and T. Miyakawa, “Impact of brain-behavior phenotypying of genetically-engineered mice on research of neuropsychiatric disorders,” Neuroscience Research, vol. 58, no. 2, pp. 124–132, 2007.
[9]  D. J. Gerber, D. Hall, T. Miyakawa et al., “Evidence for association of schizophrenia with genetic variation in the 8p21.3 gene, PPP3CC, encoding the calcineurin gamma subunit,” Proceedings of the National Academy of Sciences of the United States of America, vol. 100, no. 15, pp. 8993–8998, 2003.
[10]  T. Miyakawa, L. M. Leiter, D. J. Gerber et al., “Conditional calcineurin knockout mice exhibit multiple abnormal behaviors related to schizophrenia,” Proceedings of the National Academy of Sciences of the United States of America, vol. 100, no. 15, pp. 8987–8992, 2003.
[11]  D. G. Winder and J. D. Sweatt, “Roles of serine/threonine phosphatases in hippocampal synaptic plasticity,” Nature Reviews Neuroscience, vol. 2, no. 7, pp. 461–474, 2001.
[12]  H. Shoji, H. Hagihara, K. Takao, S. Hattori, and T. Miyakawa, “T-maze forced alternation and left-right discrimination tasks for assessing working and reference memory in mice,” Journal of Visualized Experiments, no. 60, article e3300, 2012.
[13]  N. Matsuo, N. Yamasaki, K. Ohira, et al., “Neural activity changes underlying the working memory deficit in alpha-CaMKII heterozygous knockout mice,” Frontiers in Behavioral Neuroscience, vol. 3, article 20, 2009.
[14]  C. E. Young, K. Arima, J. Xie et al., “SNAP-25 deficit and hippocampal connectivity in schizophrenia,” Cerebral Cortex, vol. 8, no. 3, pp. 261–268, 1998.
[15]  C. M. Lewis, D. F. Levinson, L. H. Wise, et al., “Genome scan meta-analysis of schizophrenia and bipolar disorder, part II: Schizophrenia,” American Journal of Human Genetics, vol. 73, no. 1, pp. 34–48, 2003.
[16]  C. L. Barr, Y. Feng, K. Wigg et al., “Identification of DNA variants in the SNAP-25 gene and linkage study of these polymorphisms and attention-deficit hyperactivity disorder,” Molecular Psychiatry, vol. 5, no. 4, pp. 405–409, 2000.
[17]  T. K. Choi, H. S. Lee, J. W. Kim et al., “Support for the MnlI polymorphism of SNAP25; a Korean ADHD case-control study,” Molecular Psychiatry, vol. 12, no. 3, pp. 224–226, 2007.
[18]  Y. Zhang, A. P. Vilaythong, D. Yoshor, and J. L. Noebels, “Elevated thalamic low-voltage-activated currents precede the onset of absence epilepsy in the SNAP25-deficient mouse mutant Coloboma,” The Journal of Neuroscience, vol. 24, no. 22, pp. 5239–5248, 2004.
[19]  M. Kataoka, R. Kuwahara, R. Matsuo, M. Sekiguchi, K. Inokuchi, and M. Takahashi, “Development- and activity-dependent regulation of SNAP-25 phosphorylation in rat brain,” Neuroscience Letters, vol. 407, no. 3, pp. 258–262, 2006.
[20]  Y. Horiuchi, H. Ishiguro, M. Koga et al., “Support for association of the PPP3CC gene with schizophrenia,” Molecular Psychiatry, vol. 12, no. 10, pp. 891–893, 2007.
[21]  H. Zeng, S. Chattarji, M. Barbarosie et al., “Forebrain-specific calcineurin knockout selectively impairs bidirectional synaptic plasticity and working/episodic-like memory,” Cell, vol. 107, no. 5, pp. 617–629, 2001.
[22]  H. Hagihara, K. Toyama, N. Yamasaki, and T. Miyakawa, “Dissection of hippocampal dentate gyrus from adult mouse,” Journal of Visualized Experiments, no. 33, 2009.
[23]  L. M. Valor, D. Jancic, R. Lujan, and A. Barco, “Ultrastructural and transcriptional profiling of neuropathological misregulation of CREB function,” Cell Death & Differentiation, vol. 17, no. 10, pp. 1636–1644, 2010.
[24]  K. Merz, S. Herold, and D. C. Lie, “CREB in adult neurogenesis—master and partner in the development of adult-born neurons?” The European Journal of Neuroscience, vol. 33, no. 6, pp. 1078–1086, 2011.
[25]  K. Ohira and T. Miyakawa, “Chronic treatment with fluoxetine for more than 6 weeks decreases neurogenesis in the subventricular zone of adult mice,” Molecular Brain, vol. 4, no. 1, article 10, 2011.
[26]  S. C. Dulawa, K. A. Holick, B. Gundersen, and R. Hen, “Effects of chronic fluoxetine in animal models of anxiety and depression,” Neuropsychopharmacology, vol. 29, no. 7, pp. 1321–1330, 2004.
[27]  R. Alonso, G. Griebel, G. Pavone, J. Stemmelin, G. Le Fur, and P. Soubrié, “Blockade of CRF1 or V1B receptors reverses stress-induced suppression of neurogenesis in a mouse model of depression,” Molecular Psychiatry, vol. 9, no. 3, 224 pages, 2004.
[28]  D. J. David, B. A. Samuels, Q. Rainer et al., “Neurogenesis-dependent and -independent effects of fluoxetine in an animal model of anxiety/depression,” Neuron, vol. 62, no. 4, pp. 479–493, 2009.
[29]  K. Kobayashi, Y. Ikeda, and H. Suzuki, “Behavioral destabilization induced by the selective serotonin reuptake inhibitor fluoxetine,” Molecular Brain, vol. 4, no. 1, article 12, 2011.
[30]  J. Detour, H. Schroeder, D. Desor, and A. Nehlig, “A 5-month period of epilepsy impairs spatial memory, decreases anxiety, but spares object recognition in the lithium-pilocarpine model in adult rats,” Epilepsia, vol. 46, no. 4, pp. 499–508, 2005.
[31]  I. Sillaber, M. Panhuysen, M. S. H. Henniger et al., “Profiling of behavioral changes and hippocampal gene expression in mice chronically treated with the SSRI paroxetine,” Psychopharmacology, vol. 200, no. 4, pp. 557–572, 2008.
[32]  H. Hagihara, K. Ohira, K. Toyama, and T. Miyakawa, “Expression of the AMPA receptor subunits GluR1 and GluR2 is associated with granule cell maturation in the dentate gyrus,” Frontiers in Neurogenesis, vol. 5, article 100, 2011.
[33]  H. Hagihara, M. Hara, K. Tsunekawa, Y. Nakagawa, M. Sawada, and K. Nakano, “Tonic-clonic seizures induce division of neuronal progenitor cells with concomitant changes in expression of neurotrophic factors in the brain of pilocarpine-treated mice,” Molecular Brain Research, vol. 139, no. 2, pp. 258–266, 2005.
[34]  O. K. Okamoto, L. Janjoppi, F. M. Bonone et al., “Whole transcriptome analysis of the hippocampus: toward a molecular portrait of epileptogenesis,” BMC Genomics, vol. 11, no. 1, article 230, 2010.
[35]  B. H. Cha, C. Akman, D. C. Silveira, X. Liu, and G. L. Holmes, “Spontaneous recurrent seizure following status epilepticus enhances dentate gyrus neurogenesis,” Brain and Development, vol. 26, no. 6, pp. 394–397, 2004.
[36]  N. N. Karpova, A. Pickenhagen, J. Lindholm, et al., “Fear erasure in mice requires synergy between antidepressant drugs and extinction training,” Science, vol. 334, no. 6063, pp. 1731–1734, 2011.
[37]  T. K. Hensch, “Critical period plasticity in local cortical circuits,” Nature Reviews Neuroscience, vol. 6, no. 11, pp. 877–888, 2005.
[38]  S. Berretta, “Extracellular matrix abnormalities in schizophrenia,” Neuropharmacology, vol. 62, no. 3, pp. 1584–1597, 2012.
[39]  T. Pizzorusso, P. Medini, N. Berardi, S. Chierzi, J. W. Fawcett, and L. Maffei, “Reactivation of ocular dominance plasticity in the adult visual cortex,” Science, vol. 298, no. 5596, pp. 1248–1251, 2002.
[40]  N. Gogolla, P. Caroni, A. Lüthi, and C. Herry, “Perineuronal nets protect fear memories from erasure,” Science, vol. 325, no. 5945, pp. 1258–1261, 2009.
[41]  J. F. M. Vetencourt, A. Sale, A. Viegi et al., “The antidepressant fluoxetine restores plasticity in the adult visual cortex,” Science, vol. 320, no. 5874, pp. 385–388, 2008.
[42]  E. S. Lein, M. J. Hawrylycz, N. Ao, et al., “Genome-wide atlas of gene expression in the adult mouse brain,” Nature, vol. 445, no. 7124, pp. 168–176, 2007.
[43]  C. A. Altar, L. W. Jurata, V. Charles et al., “Deficient hippocampal neuron expression of proteasome, ubiquitin, and mitochondrial genes in multiple schizophrenia cohorts,” Biological Psychiatry, vol. 58, no. 2, pp. 85–96, 2005.
[44]  T. M. Hyde, B. K. Lipska, T. Ali et al., “Expression of GABA signaling molecules KCC2, NKCC1, and GAD1 in cortical development and schizophrenia,” The Journal of Neuroscience, vol. 31, no. 30, pp. 11088–11095, 2011.
[45]  P. Blaesse, M. S. Airaksinen, C. Rivera, and K. Kaila, “Cation-chloride cotransporters and neuronal function,” Neuron, vol. 61, no. 6, pp. 820–838, 2009.
[46]  D. A. Lewis, D. A. Cruz, D. S. Melchitzky, and J. N. Pierri, “Lamina-specific deficits in parvalbumin-immunoreactive varicosities in the prefrontal cortex of subjects with schizophrenia: evidence for fewer projections from the thalamus,” The American Journal of Psychiatry, vol. 158, no. 9, pp. 1411–1422, 2001.
[47]  G. P. Reynolds, Z. J. Zhang, and C. L. Beasley, “Neurochemical correlates of cortical GABAergic deficits in schizophrenia: selective losses of calcium binding protein immunoreactivity,” Brain Research Bulletin, vol. 55, no. 5, pp. 579–584, 2001.
[48]  C. L. Beasley, Z. J. Zhang, I. Patten, and G. P. Reynolds, “Selective deficits in prefrontal cortical GABAergic neurons in schizophrenia defined by the presence of calcium-binding proteins,” Biological Psychiatry, vol. 52, no. 7, pp. 708–715, 2002.
[49]  C. L. Beasley and G. P. Reynolds, “Parvalbumin-immunoreactive neurons are reduced in the prefrontal cortex of schizophrenics,” Schizophrenia Research, vol. 24, no. 3, pp. 349–355, 1997.
[50]  A. Y. Wang, K. M. Lohmann, C. K. Yang, et al., “Bipolar disorder type 1 and schizophrenia are accompanied by decreased density of parvalbumin- and somatostatin-positive interneurons in the parahippocampal region,” Acta Neuropathologica, vol. 122, no. 5, pp. 615–626, 2011.
[51]  B. W. Okaty, M. N. Miller, K. Sugino, C. M. Hempel, and S. B. Nelson, “Transcriptional and electrophysiological maturation of neocortical fast-spiking GABAergic interneurons,” The Journal of Neuroscience, vol. 29, no. 21, pp. 7040–7052, 2009.
[52]  M. J. Gandal, A. M. Nesbitt, R. M. McCurdy, and M. D. Alter, “Measuring the maturity of the fast-spiking interneuron transcriptional program in autism, schizophrenia, and bipolar disorder,” PLoS ONE, vol. 7, no. 8, Article ID e41215, 2012.
[53]  H. Pantazopoulos, T. U. W. Woo, M. P. Lim, N. Lange, and S. Berretta, “Extracellular matrix-glial abnormalities in the amygdala and entorhinal cortex of subjects diagnosed with schizophrenia,” Archives of General Psychiatry, vol. 67, no. 2, pp. 155–166, 2010.
[54]  H. Pantazopoulos, N. Lange, R. J. Baldessarini, and S. Berretta, “Parvalbumin neurons in the entorhinal cortex of subjects diagnosed with bipolar disorder or schizophrenia,” Biological Psychiatry, vol. 61, no. 5, pp. 640–652, 2007.
[55]  S. Fukuda, Y. Yamasaki, T. Iwaki et al., “Characterization of the biological functions of a transcription factor, c-myc intron binding protein 1 (MIBP1),” Journal of Biochemistry, vol. 131, no. 3, pp. 349–357, 2002.
[56]  S. M. Purcell, N. R. Wray, J. L. Stone, et al., “Common polygenic variation contributes to risk of schizophrenia and bipolar disorder,” Nature, vol. 460, no. 7256, pp. 748–752, 2009.
[57]  J. Shi, D. F. Levinson, J. Duan, et al., “Common variants on chromosome 6p22. 1 are associated with schizophrenia,” Nature, vol. 460, no. 7256, pp. 753–757, 2009.
[58]  Y. Shi, Z. Li, Q. Xu, et al., “Common variants on 8p12 and 1q24. 2 confer risk of schizophrenia,” Nature Genetics, vol. 43, no. 12, pp. 1224–1227, 2011.
[59]  H. Stefansson, R. A. Ophoff, S. Steinberg, et al., “Common variants conferring risk of schizophrenia,” Nature, vol. 460, no. 7256, pp. 744–747, 2009.
[60]  W. H. Yue, H.-F. Wang, L. D. Sun, et al., “Genome-wide association study identifies a susceptibility locus for schizophrenia in Han Chinese at 11p11. 2,” Nature Genetics, vol. 43, no. 12, pp. 1228–1231, 2011.
[61]  P. J. van den Elsen, S. J. P. Gobin, M. C. J. A. van Eggermond, and A. Peijnenburg, “Regulation of MHC class I and II gene transcription: differences and similarities,” Immunogenetics, vol. 48, no. 3, pp. 208–221, 1998.
[62]  M. Y. Kimura, H. Hosokawa, M. Yamashita et al., “Regulation of T helper type 2 cell differentiation by murine Schnurri-2,” Journal of Experimental Medicine, vol. 201, no. 3, pp. 397–408, 2005.
[63]  M. Y. Kimura, C. Iwamura, A. Suzuki, et al., “Schnurri-2 controls memory Th1 and Th2 cell numbers in vivo,” Journal of Immunology, vol. 178, no. 8, pp. 4926–4936, 2007.
[64]  N. Muller and M. Schwarz, “Schizophrenia as an inflammation-mediated dysbalance of glutamatergic neurotransmission,” Neurotoxicity Research, vol. 10, no. 2, pp. 131–148, 2006.
[65]  B. H. Miller, L. E. Schultz, A. Gulati, M. D. Cameron, and M. T. Pletcher, “Genetic regulation of behavioral and neuronal responses to fluoxetine,” Neuropsychopharmacology, vol. 33, no. 6, pp. 1312–1322, 2008.
[66]  M. Kataoka, S. Yamamori, E. Suzuki, et al., “A single amino acid mutation in SNAP-25 induces anxiety-related behavior in mouse,” PLoS ONE, vol. 6, no. 9, Article ID e25158, 2011.
[67]  S. Otsuka, S. Yamamori, S. Watanabe, et al., “PKC-dependent phosphorylation of SNAP-25 plays a crucial role in the suppression of epileptogenesis and anxiety-related behavior in postnatal period of mouse,” Neuroscience Research, vol. 71, supplement, article e296, 2011.
[68]  A. Vezzani, “Inflammation and epilepsy,” Epilepsy Currents, vol. 5, no. 1, pp. 1–6, 2005.
[69]  A. Vezzani and T. Granata, “Brain inflammation in epilepsy: experimental and clinical evidence,” Epilepsia, vol. 46, no. 11, pp. 1724–1743, 2005.
[70]  O. Tomkins, O. Friedman, S. Ivens et al., “Blood-brain barrier disruption results in delayed functional and structural alterations in the rat neocortex,” Neurobiology of Disease, vol. 25, no. 2, pp. 367–377, 2007.
[71]  E. A. van Vliet, S. da C. Araújo, S. Redeker, R. Van Schaik, E. Aronica, and J. A. Gorter, “Blood-brain barrier leakage may lead to progression of temporal lobe epilepsy,” Brain, vol. 130, no. 2, pp. 521–534, 2007.
[72]  L. Uva, L. Librizzi, N. Marchi et al., “Acute induction of epileptiform discharges by pilocarpine in the in vitro isolated guinea-pig brain requires enhancement of blood-brain barrier permeability,” Neuroscience, vol. 151, no. 1, pp. 303–312, 2008.
[73]  P. F. Fabene, G. N. Mora, M. Martinello, et al., “A role for leukocyte-endothelial adhesion mechanisms in epilepsy,” Nature Medicine, vol. 14, no. 12, pp. 1377–1383, 2008.
[74]  M. Zhou, W. Li, S. Huang, et al., “mTOR inhibition ameliorates cognitive and affective deficits caused by Disc1 knockdown in adult-born dentate granule neurons,” Neuron, vol. 77, no. 4, pp. 647–654, 2013.
[75]  D. St Clair, D. Blackwood, W. Muir et al., “Association within a family of a balanced autosomal translocation with major mental illness,” The Lancet, vol. 336, no. 8706, pp. 13–16, 1990.
[76]  N. Burnashev, H. Monyer, P. H. Seeburg, and B. Sakmann, “Divalent ion permeability of AMPA receptor channels is dominated by the edited form of a single subunit,” Neuron, vol. 8, no. 1, pp. 189–198, 1992.
[77]  M. Hollmann, M. Hartley, and S. Heinemann, “Ca2+ permeability of KA-AMPA-gated glutamate receptor channels depends on subunit composition,” Science, vol. 252, no. 5007, pp. 851–853, 1991.
[78]  R. I. Hume, R. Dingledine, and S. F. Heinemann, “Identification of a site in glutamate receptor subunits that controls calcium permeability,” Science, vol. 253, no. 5023, pp. 1028–1031, 1991.
[79]  C. Zhao, W. Deng, and F. H. Gage, “Mechanisms and functional implications of adult neurogenesis,” Cell, vol. 132, no. 4, pp. 645–660, 2008.
[80]  S. Jessberger and G. Kempermann, “Adult-born hippocampal neurons mature into activity-dependent responsiveness,” The European Journal of Neuroscience, vol. 18, no. 10, pp. 2707–2712, 2003.
[81]  P. S. Goldman-Rakic, “Working memory dysfunction in schizophrenia,” Journal of Neuropsychiatry and Clinical Neurosciences, vol. 6, no. 4, pp. 348–357, 1994.
[82]  B. Elvev?g and T. E. Goldberg, “Cognitive impairment in schizophrenia is the core of the disorder,” Critical Reviews in Neurobiology, vol. 14, no. 1, pp. 1–21, 2000.
[83]  D. L. Braff and M. A. Geyer, “Sensorimotor gating and schizophrenia: human and animal model studies,” Archives of General Psychiatry, vol. 47, no. 2, pp. 181–188, 1990.
[84]  American Psychiatric Association, Diagnostic and Statistical Manual of Mental Disorders, American Psychiatric Association, Washington, DC, USA, 4th edition, 1994.
[85]  R. R. Gainetdinov, A. R. Mohn, L. M. Bohn, and M. G. Caron, “Glutamatergic modulation of hyperactivity in mice lacking the dopamine transporter,” Proceedings of the National Academy of Sciences of the United States of America, vol. 98, no. 20, pp. 11047–11054, 2001.
[86]  P. R. Maycox, F. Kelly, A. Taylor et al., “Analysis of gene expression in two large schizophrenia cohorts identifies multiple changes associated with nerve terminal function,” Molecular Psychiatry, vol. 14, no. 12, pp. 1083–1094, 2009.
[87]  Z. J. Zhang and G. P. Reynolds, “A selective decrease in the relative density of parvalbumin-immunoreactive neurons in the hippocampus in schizophrenia,” Schizophrenia Research, vol. 55, no. 1-2, pp. 1–10, 2002.
[88]  M. B. Knable, B. M. Barci, M. J. Webster, J. Meador-Woodruff, and E. F. Torrey, “Molecular abnormalities of the hippocampus in severe psychiatric illness: postmortem findings from the Stanley Neuropathology Consortium,” Molecular Psychiatry, vol. 9, no. 6, pp. 609–620, 2004.
[89]  F. M. Benes, B. Lim, D. Matzilevich, J. P. Walsh, S. Subburaju, and M. Minns, “Regulation of the GABA cell phenotype in hippocampus of schizophrenics and bipolars,” Proceedings of the National Academy of Sciences of the United States of America, vol. 104, no. 24, pp. 10164–10169, 2007.
[90]  J. N. Pierri, A. S. Chaudry, T. U. W. Woo, and D. A. Lewis, “Alterations in chandelier neuron axon terminals in the prefrontal cortex of schizophrenic subjects,” The American Journal of Psychiatry, vol. 156, no. 11, pp. 1709–1719, 1999.
[91]  J. Gallinat, G. Winterer, C. S. Herrmann, and D. Senkowski, “Reduced oscillatory gamma-band responses in unmedicated schizophrenic patients indicate impaired frontal network processing,” Clinical Neurophysiology, vol. 115, no. 8, pp. 1863–1874, 2004.
[92]  L. V. Moran and L. E. Hong, “High vs low frequency neural oscillations in schizophrenia,” Schizophrenia Bulletin, vol. 37, no. 4, pp. 659–663, 2011.
[93]  S. R. Sponheim, B. A. Clementz, W. G. Iacono, and M. Beiser, “Resting EEG in first-episode and chronic schizophrenia,” Psychophysiology, vol. 31, no. 1, pp. 37–43, 1994.
[94]  A. Abi-Dargham, “Do we still believe in the dopamine hypothesis? New data bring new evidence,” The International Journal of Neuropsychopharmacology, vol. 7, supplement 1, pp. S1–S5, 2004.
[95]  G. Winterer and D. R. Weinberger, “Genes, dopamine and cortical signal-to-noise ratio in schizophrenia,” Trends in Neuroscience, vol. 27, no. 11, pp. 683–690, 2004.
[96]  D. A. Cousins, K. Butts, and A. H. Young, “The role of dopamine in bipolar disorder,” Bipolar Disorders, vol. 11, no. 8, pp. 787–806, 2009.
[97]  G. Novak and P. Seeman, “Hyperactive mice show elevated D2High receptors, a model for schizophrenia: calcium/calmodulin-dependent kinase II alpha knockouts,” Synapse, vol. 64, no. 10, pp. 794–800, 2010.
[98]  C. R. Maxwell, S. J. Kanes, T. Abel, and S. J. Siegel, “Phosphodiesterase inhibitors: a novel mechanism for receptor-independent antipsychotic medications,” Neuroscience, vol. 129, no. 1, pp. 101–107, 2004.
[99]  S. J. Kanes, J. Tokarczyk, S. J. Siegel, W. Bilker, T. Abel, and M. P. Kelly, “Rolipram: a specific phosphodiesterase 4 inhibitor with potential antipsychotic activity,” Neuroscience, vol. 144, no. 1, pp. 239–246, 2007.
[100]  J. Lisman, H. Schulman, and H. Cline, “The molecular basis of CaMKII function in synaptic and behavioural memory,” Nature Reviews Neuroscience, vol. 3, no. 3, pp. 175–190, 2002.
[101]  D. R. Weinberger, “Implications of normal brain development for the pathogenesis of schizophrenia,” Archives of General Psychiatry, vol. 44, no. 7, pp. 660–669, 1987.
[102]  D. R. Weinberger and R. K. McClure, “Neurotoxicity, neuroplasticity, and magnetic resonance imaging morphometry: what is happening in the schizophrenic brain?” Archives of General Psychiatry, vol. 59, no. 6, pp. 553–558, 2002.
[103]  S. Marenco and D. R. Weinberger, “The neurodevelopmental hypothesis of schizophrenia: following a trail of evidence from cradle to grave,” Development and Psychopathology, vol. 12, no. 3, pp. 501–527, 2000.
[104]  T. D. Cannon, I. M. Rosso, C. E. Bearden, L. E. Sanchez, and T. Hadley, “A prospective cohort study of neurodevelopmental processes in the genesis and epigenesis of schizophrenia,” Development and Psychopathology, vol. 11, no. 3, pp. 467–485, 1999.
[105]  B. K. Lipska, “Using animal models to test a neurodevelopmental hypothesis of schizophrenia,” Journal of Psychiatry and Neuroscience, vol. 29, no. 4, pp. 282–286, 2004.
[106]  C. T. Ekdahl, J.-H. Claasen, S. Bonde, Z. Kokaia, and O. Lindvall, “Inflammation is detrimental for neurogenesis in adult brain,” Proceedings of the National Academy of Sciences of the United States of America, vol. 100, no. 23, pp. 13632–13637, 2003.
[107]  S. Das and A. Basu, “Inflammation: a new candidate in modulating adult neurogenesis,” Journal of Neuroscience Research, vol. 86, no. 6, pp. 1199–1208, 2008.
[108]  R. L. Hunter, N. Dragicevic, K. Seifert et al., “Inflammation induces mitochondrial dysfunction and dopaminergic neurodegeneration in the nigrostriatal system,” Journal of Neurochemistry, vol. 100, no. 5, pp. 1375–1386, 2007.
[109]  M. T. Fischer, R. Sharma, J. L. Lim, et al., “NADPH oxidase expression in active multiple sclerosis lesions in relation to oxidative tissue damage and mitochondrial injury,” Brain, vol. 135, no. 3, pp. 886–899, 2012.
[110]  Y. Pang, Z. Cai, and P. G. Rhodes, “Disturbance of oligodendrocyte development, hypomyelination and white matter injury in the neonatal rat brain after intracerebral injection of lipopolysaccharide,” Developmental Brain Research, vol. 140, no. 2, pp. 205–214, 2003.
[111]  W. Bruck, R. Pfortner, T. Pham, et al., “Reduced astrocytic NF-κB activation by laquinimod protects from cuprizone-induced demyelination,” Acta Neuropathologica, vol. 124, no. 3, pp. 411–424, 2012.
[112]  A. Reif, S. Fritzen, M. Finger et al., “Neural stem cell proliferation is decreased in schizophrenia, but not in depression,” Molecular Psychiatry, vol. 11, no. 5, pp. 514–522, 2006.
[113]  M. S. Keshavan, H. A. Nasrallah, and R. Tandon, “Schizophrenia, “Just the Facts” 6. Moving ahead with the schizophrenia concept: from the elephant to the mouse,” Schizophrenia Research, vol. 127, no. 1–3, pp. 3–13, 2011.
[114]  H. Nawa and N. Takei, “Recent progress in animal modeling of immune inflammatory processes in schizophrenia: implication of specific cytokines,” Neuroscience Research, vol. 56, no. 1, pp. 2–13, 2006.
[115]  P. H. Patterson, “Immune involvement in schizophrenia and autism: etiology, pathology and animal models,” Behavioural Brain Research, vol. 204, no. 2, pp. 313–321, 2009.
[116]  A. S. Brown and P. H. Patterson, “Maternal infection and schizophrenia: implications for prevention,” Schizophrenia Bulletin, vol. 37, no. 2, pp. 284–290, 2011.
[117]  E. Y. Hsiao, S. W. McBride, J. Chow, S. K. Mazmanian, and P. H. Patterson, “Modeling an autism risk factor in mice leads to permanent immune dysregulation,” Proceedings of the National Academy of Sciences of the United States of America, vol. 109, no. 31, pp. 12776–12781, 2012.
[118]  P. H. Patterson, “Maternal infection: window on neuroimmune interactions in fetal brain development and mental illness,” Current Opinion in Neurobiology, vol. 12, no. 1, pp. 115–118, 2002.
[119]  L. Shi, S. H. Fatemi, R. W. Sidwell, and P. H. Patterson, “Maternal influenza infection causes marked behavioral and pharmacological changes in the offspring,” The Journal of Neuroscience, vol. 23, no. 1, pp. 297–302, 2003.
[120]  J. P. A. Ioannidis, E. E. Ntzani, T. A. Trikalinos, and D. G. Contopoulos-Ioannidis, “Replication validity of genetic association studies,” Nature Genetics, vol. 29, no. 3, pp. 306–309, 2001.
[121]  M. Ayalew, H. Le-Niculescu, D. F. Levey, et al., “Convergent functional genomics of schizophrenia: from comprehensive understanding to genetic risk prediction,” Molecular Psychiatry, vol. 17, no. 9, pp. 887–905, 2012.
[122]  K. V. Chowdari, K. Mirnics, P. Semwal et al., “Association and linkage analyses of RGS4 polymorphisms in schizophrenia,” Human Molecular Genetics, vol. 11, no. 12, pp. 1373–1380, 2002.
[123]  K. Mirnics, F. A. Middleton, G. D. Stanwood, D. A. Lewis, and P. Levitt, “Disease-specific changes in regulator of G-protein signaling 4 (RGS4) expression in schizophrenia,” Molecular Psychiatry, vol. 6, no. 3, pp. 293–301, 2001.
[124]  T. L. Petryshen, F. A. Middleton, A. R. Tahl, et al., “Genetic investigation of chromosome 5q GABAA receptor subunit genes in schizophrenia,” Molecular Psychiatry, vol. 10, no. 12, pp. 1074–1088, 2005.
[125]  N. C. Allen, S. Bagade, M. B. McQueen et al., “Systematic meta-analyses and field synopsis of genetic association studies in schizophrenia: the SzGene database,” Nature Genetics, vol. 40, no. 7, pp. 827–834, 2008.
[126]  S. J. Huffaker, J. Chen, K. K. Nicodemus et al., “A primate-specific, brain isoform of KCNH2 affects cortical physiology, cognition, neuronal repolarization and risk of schizophrenia,” Nature Medicine, vol. 15, no. 5, pp. 509–518, 2009.
[127]  R. Zakharyan, A. Khoyetsyan, A. Arakelyan, et al., “Association of C1QB gene polymorphism with schizophrenia in Armenian population,” BMC Medical Genetics, vol. 12, article 126, 2011.
[128]  J. Ekelund, D. Lichtermann, I. Hovatta et al., “Genome-wide scan for schizophrenia in the Finnish population: evidence for a locus on chromosome 7q22,” Human Molecular Genetics, vol. 9, no. 7, pp. 1049–1057, 2000.
[129]  W. Yan, X.-Y. Guan, E. D. Green et al., “Childhood-onset schizophrenia/autistic disorder and t(1;7) reciprocal translocation: identification of a BAC contig spanning the translocation breakpoint at 7q21,” American Journal of Medical Genetics, vol. 96, no. 6, pp. 749–753, 2000.
[130]  M. Bradford, M. H. Law, A. D. Stewart, D. J. Shaw, I. L. Megson, and J. Wei, “The TGM2 gene is associated with schizophrenia in a british population,” American Journal of Medical Genetics B, vol. 150, no. 3, pp. 335–340, 2009.
[131]  S. W. Flynn, D. J. Lang, A. L. Mackay et al., “Abnormalities of myelination in schizophrenia detected in vivo with MRI, and post-mortem with analysis of oligodendrocyte proteins,” Molecular Psychiatry, vol. 8, no. 9, pp. 811–820, 2003.
[132]  K. Iwamoto, M. Bundo, and T. Kato, “Altered expression of mitochondria-related genes in postmortem brains of patients with bipolar disorder or schizophrenia, as revealed by large-scale DNA microarray analysis,” Human Molecular Genetics, vol. 14, no. 2, pp. 241–253, 2005.
[133]  D. S. Olton and B. C. Papas, “Spatial memory and hippocampal function,” Neuropsychologia, vol. 17, no. 6, pp. 669–682, 1979.
[134]  P. S. Goldman-Rakic, “Regional and cellular fractionation of working memory,” Proceedings of the National Academy of Sciences of the United States of America, vol. 93, no. 24, pp. 13473–13480, 1996.
[135]  R. L. McLamb, W. R. Mundy, and H. A. Tilson, “Intradentate colchicine disrupts the acquisition and performance of a working memory task in the radial arm maze,” NeuroToxicology, vol. 9, no. 3, pp. 521–528, 1988.
[136]  D. F. Emerich and T. J. Walsh, “Selective working memory impairments following intradentate injection of colchicine: attenuation of the behavioral but not the neuropathological effects by gangliosides GM1 and AGF2,” Physiology & Behavior, vol. 45, no. 1, pp. 93–101, 1989.
[137]  A. M. Morris, J. C. Churchwell, R. P. Kesner, and P. E. Gilbert, “Selective lesions of the dentate gyrus produce disruptions in place learning for adjacent spatial locations,” Neurobiology of Learning and Memory, vol. 97, no. 3, pp. 326–331, 2012.
[138]  R. P. Kesner, “Behavioral functions of the CA3 subregion of the hippocampus,” Learning and Memory, vol. 14, no. 11, pp. 771–781, 2007.
[139]  M. L. Shapiro and D. S. Olton, Memory Systems 1994, MIT Press, 1994.
[140]  J. B. Aimone, J. Wiles, and F. H. Gage, “Potential role for adult neurogenesis in the encoding of time in new memories,” Nature Neuroscience, vol. 9, no. 6, pp. 723–727, 2006.
[141]  M. A. Yassa and C. E. L. Stark, “Pattern separation in the hippocampus,” Trends in Neurosciences, vol. 34, no. 10, pp. 515–525, 2011.
[142]  D. Marr, “Simple memory: a theory for archicortex,” Philosophical Transactions of the Royal Society of London B, vol. 262, no. 841, pp. 23–81, 1971.
[143]  B. L. McNaughton and R. G. M. Morris, “Hippocampal synaptic enhancement and information storage within a distributed memory system,” Trends in Neurosciences, vol. 10, no. 10, pp. 408–415, 1987.
[144]  R. C. O'Reilly and J. L. McClelland, “Hippocampal conjunctive encoding, storage, and recall: avoiding a trade-off,” Hippocampus, vol. 4, no. 6, pp. 661–682, 1994.
[145]  A. Treves, A. Tashiro, M. E. Witter, and E. I. Moser, “What is the mammalian dentate gyrus good for?” Neuroscience, vol. 154, no. 4, pp. 1155–1172, 2008.
[146]  P. E. Gilbert, R. P. Kesner, and I. Lee, “Dissociating hippocampal subregions: a double dissociation between dentate gyrus and CA1,” Hippocampus, vol. 11, no. 6, pp. 626–636, 2001.
[147]  P. E. Gilbert and R. P. Kesner, “Localization of function within the dorsal hippocampus: the role of the CA3 subregion in paired-associate learning,” Behavioral Neuroscience, vol. 117, no. 6, pp. 1385–1394, 2003.
[148]  I. Lee and R. P. Kesner, “Different contributions of dorsal hippocampal subregios to emory acquisation and retrieval in contextual fear-conditioning,” Hippocampus, vol. 14, no. 3, pp. 301–310, 2004.
[149]  I. Lee and R. P. Kesner, “Encoding versus retrieval of spatial memory: double dissociation between the dentate gyrus and the perforant path inputs into CA3 in the dorsal hippocampus,” Hippocampus, vol. 14, no. 1, pp. 66–76, 2004.
[150]  J. K. Leutgeb, S. Leutgeb, M.-B. Moser, and E. I. Moser, “Pattern separation in the dentate gyrus and CA3 of the hippocampus,” Science, vol. 315, no. 5814, pp. 961–966, 2007.
[151]  T. J. McHugh, M. W. Jones, J. J. Quinn et al., “Dentate gyrus NMDA receptors mediate rapid pattern separation in the hippocampal network,” Science, vol. 317, no. 5834, pp. 94–99, 2007.
[152]  J. B. Aimone, W. Deng, and F. H. Gage, “Resolving new memories: a critical look at the dentate gyrus, adult neurogenesis, and pattern separation,” Neuron, vol. 70, no. 4, pp. 589–596, 2011.
[153]  A. Marín-Burgin, L. A. Mongiat, M. B. Pardi, and A. F. Schinder, “Unique processing during a period of high excitation/inhibition balance in adult-born neurons,” Science, vol. 335, no. 6073, pp. 1238–1242, 2012.
[154]  T. Nakashiba, J. D. Cushman, K. A. Pelkey, et al., “Young dentate granule cells mediate pattern separation, whereas old granule cells facilitate pattern completion,” Cell, vol. 149, no. 1, pp. 188–201, 2012.
[155]  W. A. M. Swinkels, J. Kuyk, R. van Dyck, and P. Spinhoven, “Psychiatric comorbidity in epilepsy,” Epilepsy and Behavior, vol. 7, no. 1, pp. 37–50, 2005.
[156]  P. Cifelli and A. A. Grace, “Pilocarpine-induced temporal lobe epilepsy in the rat is associated with increased dopamine neuron activity,” The International Journal of Neuropsychopharmacology, vol. 15, no. 7, pp. 957–964, 2012.
[157]  D. J. Lodge and A. A. Grace, “Hippocampal dysregulation of dopamine system function and the pathophysiology of schizophrenia,” Trends in Pharmacological Sciences, vol. 32, no. 9, pp. 507–513, 2011.
[158]  C. A. Tamminga, S. Southcott, C. Sacco, A. D. Wagner, and S. Ghose, “Glutamate dysfunction in hippocampus: relevance of dentate gyrus and CA3 signaling,” Schizophrenia Bulletin, vol. 38, no. 5, pp. 927–935, 2012.
[159]  C. O. Lacefield, V. Itskov, T. Reardon, R. Hen, and J. A. Gordon, “Effects of adult-generated granule cells on coordinated network activity in the dentate gyrus,” Hippocampus, vol. 22, no. 1, pp. 106–116, 2012.
[160]  D. A. Henze, N. N. Urban, and G. Barrionuevo, “The multifarious hippocampal mossy fiber pathway: a review,” Neuroscience, vol. 98, no. 3, pp. 407–427, 2000.
[161]  J. Song, K. M. Christian, G. Ming, and H. Song, “Modification of hippocampal circuitry by adult neurogenesis,” Developmental Neurobiology, vol. 72, no. 7, pp. 1032–1043, 2012.
[162]  A. A. Grace, S. B. Floresco, Y. Goto, and D. J. Lodge, “Regulation of firing of dopaminergic neurons and control of goal-directed behaviors,” Trends in Neurosciences, vol. 30, no. 5, pp. 220–227, 2007.
[163]  I. I. Gottesman and T. D. Gould, “The endophenotype concept in psychiatry: etymology and strategic intentions,” The American Journal of Psychiatry, vol. 160, no. 4, pp. 636–645, 2003.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133