Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-19T19:40:49.769Z Has data issue: false hasContentIssue false

4 - Neuronal cell death in human neurodegenerative diseases and their animal/cell models

Published online by Cambridge University Press:  03 March 2010

Lee J. Martin
Affiliation:
Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA
Zhiping Liu
Affiliation:
Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA
Juan Troncoso
Affiliation:
Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA
Donald L. Price
Affiliation:
Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA
Martin Holcik
Affiliation:
University of Ottawa
Eric C. LaCasse
Affiliation:
University of Ottawa
Alex E. MacKenzie
Affiliation:
University of Ottawa
Robert G. Korneluk
Affiliation:
University of Ottawa
Get access

Summary

Introduction

Cell death is important in the normal histogenesis of organs, steady state kinetics of healthy adult tissues, pathogenesis of tissue damage and disease, and disease therapy. Pathologists conceived the concept of cell death as a mechanism for disease to aid in diagnosis and therapy (Virchow, 1858). Developmental biologists then realized the essential role of cell death in tissue and organ development (Glücksmann, 1951; Lockshin and Williams, 1964; Saunders, 1966). This idea was at first received with skepticism, but now it is dogma that cell number in tissues is controlled precisely in developing and adult tissues. The absence of this precise control of cell number in tissues causes cancer (impaired apoptosis is a central step toward neoplasia). Pathologic stimuli can be extrinsic or intrinsic and can inactivate normal cell death networks or can cause abrupt or delayed cell death. Cell demise can occur as multiple types of death (Schweichel and Merker, 1973; Lockshin and Zakeri, 2002). It is compelling that a goal of human disease therapy is, on the one hand, to prevent cell death in neurologic disease and, on the other hand, to stimulate cell death in malignancy. Thus, the study of cell death is fundamental to human pathobiology and disease treatment. In this chapter, recent critical views on the contributions of the different forms of cell death to human neurodegenerative diseases and their animal and cell models will be presented.

Type
Chapter
Information
Apoptosis in Health and Disease
Clinical and Therapeutic Aspects
, pp. 96 - 155
Publisher: Cambridge University Press
Print publication year: 2005

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adamec, E., Vonsattel, J. P., and Nixon, R. A. (1999). DNA strand breaks in Alzheimer's disease. Brain Res., 849, 67–77CrossRefGoogle ScholarPubMed
Al-Abdulla, N. A. and Martin, L. J. (1998). Apoptosis of retrogradely degenerating neurons occurs in association with the accumulation of perikaryal mitochondria and oxidative damage to the nucleus. Am. J. Pathol., 153, 447–56CrossRefGoogle ScholarPubMed
Al-Abdulla, N. A., Portera-Cailliau, C., and Martin, L. J. (1998). Occipital cortex ablation in adult rat causes retrograde neuronal death in the lateral geniculate nucleus that resembles apoptosis. Neuroscience, 86, 191–209CrossRefGoogle ScholarPubMed
Aloyz, R. S., Bamji, S. X., Pozniak, C. D.et al. (1998). p53 is essential for developmental neuron death regulated by the TrkA and p75 neurotrophin receptors. J. Cell Biol., 143, 1691–703CrossRefGoogle ScholarPubMed
Alvarez, A., Munoz, J. P., and Maccioni, R. B. (2001). A cdk5-p35 stable complex is involved in the β-amyloid-induced deregulation of cdk5 activity in hippocampal neurons. Exp. Cell Res., 264, 266–74CrossRefGoogle ScholarPubMed
Alves da Costa, C., Paitel, E., Vincent, B., and Checler, F. (2002). Alpha-synuclein lowers p53-dependent apoptotic of neuronal cell: abolishment by 6-hydroxydopamine and implication for Parkinson's disease. J. Biol. Chem., 277, 50980–4CrossRefGoogle ScholarPubMed
Ameisen, J. C. (2002). On the origin, evolution, and nature of programmed cell death: a timeline of four billion years. Cell Death Diff., 9, 367–93CrossRefGoogle ScholarPubMed
Amin, F., Bowen, I. D., Szegedi, Z., Mihalik, R., and Szende, B. (2000). Apoptotic and non-apoptotic modes of programmed cell death in MCF-7 human breast carcinoma cells. Cell Biol. Intl., 24, 253–60CrossRefGoogle ScholarPubMed
Anderson, A. J., Su, J. H., and Cotman, C. W. (1996). DNA damage and apoptosis in Alzheimer's disease: colocalization with c-jun immunoreactivity, relationship to brain area, and effect of postmortem delay. J. Neurosci., 16, 1710–19CrossRefGoogle ScholarPubMed
Anglade, P., Vyas, S., Javoy-Agid, F.et al. (1997). Apoptosis and autophagy in nigral neurons of patients with Parkinson's disease. Histol. Histopathol., 12, 25–31Google ScholarPubMed
Ankarcrona, M., Dypbukt, J. M., Bonfoco, E.et al. (1995). Glutamate-induced neuronal death: a succession of necrosis or apoptosis depending on mitochondrial function. Neuron, 15, 961–73CrossRefGoogle ScholarPubMed
Antonsson, B., Conti, F., Ciavatta, A.et al. (1997). Inhibition of bax channel-forming activity by bcl-2. Science, 277, 370–2CrossRefGoogle ScholarPubMed
Barron, K. D., Means, E. D., and Larsen, E . (1973). Ultrastructure of retrograde degeneration in thalamus of rat. I. Neuronal somata and dendrites. J. Neuropathol. Exp. Neurol., 32, 218–44CrossRefGoogle ScholarPubMed
Beckman, J. S., Carson, M., Smith, C. D., and Koppenol, W. H. (1993). ALS, SOD and peroxynitrite. Nature, 364, 548CrossRefGoogle ScholarPubMed
Beckman, J. S., Estōvez, A. G., Crow, J. P., and Barbeito, , L. (2001). Superoxide dismutase and the death of motoneurons in ALS. Trends Neurosci., 24 (suppl), S15–S20CrossRefGoogle ScholarPubMed
Behl, C., Davis, J. B., Klier, F. G., and Schubert, D. (1994). Amyloid beta peptide induces necrosis rather than apoptosis. Brain Res., 645, 253–64CrossRefGoogle ScholarPubMed
Bendotti, C., Calvaresi, N., Chiveri, L.et al. (2001). Early vacuolization and mitochondrial damage in motor neurons of FALS mice are not associated with apoptosis or with changes in cytochrome oxidase histochemical reactivity. J. Neurol. Sci., 191, 25–33CrossRefGoogle ScholarPubMed
Beresford, P. J., Zhang, D., Oh, D. Y.et al. (2001). Granzyme A activates an endoplasmic reticulum-associated caspase-independent nuclease to induce single-stranded DNA nicks. J. Biol. Chem., 276, 43285–93CrossRefGoogle ScholarPubMed
Bilsland, J., Roy, S., Xanthoudakis, S.et al. (2002). Caspase inhibitors attenuate 1-methyl-4-phenylpyridinum toxicity in primary cultures of mesencephalic dopaminergic neurons. J. Neurosci., 22, 2637–49CrossRefGoogle Scholar
Blum, D., Wu, Y., Nissou, M-F., Arnaud, S., Benabid, A.-L., and Verna, J.-M. (1997). p53 and Bax activation in 6-hydroxydopamine-induced apoptosis in PC12 cells. Brain Res., 751, 139–42CrossRefGoogle ScholarPubMed
Bonfanti, L., Strettoi, E., Chierzi, S.et al. (1996). Protection of retinal ganglion cells from natural and axotomy-induced cell death in neonatal transgenic mice overexpressing bcl-2. J. Neurosci., 16, 4186–94CrossRefGoogle ScholarPubMed
Bonfoco, E., Krainc, D., Ankarcrona, M., Nicotera, P., and Lipton, S. A. (1995). Apoptosis and necrosis: two distinct events induced, respectively, by mild and intense insults with N-methyl-D-aspartate or nitric oxide/superoxide in cortical cell culture. Proc. Natl. Acad. Sci. USA, 92, 7162–6CrossRefGoogle ScholarPubMed
Bradley, W. G. and Krasin, F. (1982). A new hypothesis of the etiology of amyotrophic lateral sclerosis. The DNA hypothesis. Arch. Neurol., 39, 677–80CrossRefGoogle ScholarPubMed
Bucciantini, M., Giannoni, E., Chiti, F.et al. (2002). Inherent toxicity of aggregates implies a common mechanism for misfolding diseases. Nature, 416, 507–11CrossRefGoogle ScholarPubMed
Bursch, W. (2001). The autophagosomal-lysosomal compartment in programmed cell death. Cell Death Diff., 8, 569–81CrossRefGoogle ScholarPubMed
Bursch, W., Paffe, S., Putz, B., Barthel, G., and Schulte-Hermann, R. (1990). Determination of the length of the histological stages of apoptosis in normal liver and in altered hepatic foci of rats. Carcinogenesis, 11, 847–53CrossRefGoogle ScholarPubMed
Bursztajn, S., DeSouza, R., McPhie, D. L.et al. (1998). Overexpression in neurons of human presenilin-1 or a presenilin-1 familial Alzheimer disease mutant does not enhance apoptosis. J. Neurosci., 18, 9790–9CrossRefGoogle Scholar
Calhoun, M. E., Wiederhold, K. H., Abramowski, D.et al. (1998). Neuron loss in APP transgenic mice. Nature, 395, 755–6CrossRefGoogle ScholarPubMed
Campion, D., Flaman, J. M., Brice, A.et al. (1995). Mutations of the presenilin 1 gene in families with early-onset Alzheimer's disease. Hum. Mol. Genet., 4, 2373–7CrossRefGoogle ScholarPubMed
Cardone, M. H., Roy, N., Stennicke, H. R.et al. (1998). Regulation of cell death protease caspase-9 by phosphorylation. Science, 282, 1318–21CrossRefGoogle ScholarPubMed
Cardone, M. H., Salvesen, G. S., Widmann, C., Johnson, G., and Frisch, S. M. (1997). The regulation of anoikis: MEKK-1 activation requires cleavage by caspases. Cell, 90, 315–23CrossRefGoogle ScholarPubMed
Casaccia-Bonnefil, P., Carter, B. D., Dobrowsky, R. T., and Chao, M. V. (1996). Death of oligodendrocytes mediated by the interaction of nerve growth factor with its receptor p75. Nature, 383, 716–19CrossRefGoogle ScholarPubMed
Casley, C. S., Land, J. M., Sharpe, M. A., Clark, J. B., Duchen, M. R., and Canevari, L. (2002). β-amyloid fragment 25–35 causes mitochondrial dysfunction in primary cortical neurons. Neurobiol. Dis., 10, 258–67CrossRefGoogle ScholarPubMed
Cataldo, A. M., Hamilton, D. J., and Nixon, R. A. (1994). Lysosomal abnormalities in degenerating neurons link neuronal compromise to senile plaque development in Alzheimer's disease. Brain Res., 640, 68–80CrossRefGoogle Scholar
Chan, S. L., Culmsee, C., Haughey, N., Klapper, W., and Mattson, M. P. (2002). Presenilin-1 mutations sensitize neurons to DNA damage-induced death by a mechanism involving perturbed calcium homeostasis and activation of calpains and caspase-12. Neurobiol. Dis., 11, 2–19CrossRefGoogle ScholarPubMed
Chang, Y.-C., Lee, Y.-S., Tejima, T.et al. (1998). Mdm-2 and bax, downstream mediators of the p53 response, are degraded by the ubiquitin-proteasome pathway. Cell Growth Diff., 9, 79–84Google Scholar
Chartier-Harlin, M.-C., Crawford, F., Houlden, H.et al. (1991). Early-onset Alzheimer's disease caused by mutations at codon 717 of the β-amyloid precursor protein gene. Nature, 353, 844–6CrossRefGoogle ScholarPubMed
Choi, D. W. (1992). Excitotoxic cell death. J. Neurobiol., 23, 1261–76CrossRefGoogle ScholarPubMed
Chu-Wang, I.-W. and Oppenheim, R. W. (1978). Cell death of motoneurons in the chick embryo spinal cord. I. A light and electron microscopic study of naturally occurring and induced cell loss during development. J. Comp. Neurol., 177, 33–58CrossRefGoogle ScholarPubMed
Clarke, P. G. H. (1990). Developmental cell death: morphological diversity and multiple mechanisms. Anat. Embryol., 181, 195–213CrossRefGoogle ScholarPubMed
Cory, S. and Adams, J. M. (2002). The bcl-2 family: regulators of the cellular life-or-death switch. Nat. Rev., 2, 647–56CrossRefGoogle ScholarPubMed
Crocker, S. J., Wigle, N., Liston, P.et al. (2001). NAIP protects the nigrostriatal dopamine pathway in an intrastriatal 6-OHDA rat model of Parkinson's disease. Eur. J. Neurosci., 14, 391–400CrossRefGoogle Scholar
Crompton, M. (1999). The mitochondrial permeability transition pore and its role in cell death. Biochem. J., 341, 233–49CrossRefGoogle ScholarPubMed
Cruz, J. C., Tseng, H. C., Goldman, J. A., Shih, H., and Tsai, L.-H. (2003). Aberrant Cdk5 activation by p25 triggers pathological events leading to neurodegeneration and neurofibrillary tangles. Neuron, 40, 471–83CrossRefGoogle ScholarPubMed
Culmsee, C., Zhu, X., Yu, Q. S.et al. (2001). A synthetic inhibitor of p53 protects neurons against death induced by ischemic and excitotoxic insults, and amyloid beta-peptide. J. Neurochem., 77, 220–8CrossRefGoogle ScholarPubMed
Danial, N. N., Gramm, C. F., Scorrano, L.et al. (2003). Bad and glucokinase reside in a mitochondrial complex that integrates glycolysis and apoptosis. Nature, 424, 952–6CrossRefGoogle Scholar
Dargusch, R., Piasecki, D., Tan, S., Liu, Y., and Schubert, D. (2001). The role of Bax in glutamate-induced nerve cell death. J. Neurochem., 76, 295–301CrossRefGoogle ScholarPubMed
Datta, S. R., Dudek, H., Tao, X.et al. (1997). Akt phosphorylation of Bad couples survival signals to the cell-intrinsic death machinery. Cell, 91, 231–41CrossRefGoogle ScholarPubMed
Davies, S. W., Turmaine, M., Cozens, B. A.et al. (1997). Formation of neuronal intranuclear inclusions underlies the neurological dysfunction in mice transgenic for the HD mutation. Cell, 90, 537–48CrossRefGoogle ScholarPubMed
Aguilar, J. L. G., Gorden, J. W., Rene, F.et al. (2000). Alteration of the cl-x/Bax ratio in a transgenic mouse model of amyotrophic lateral sclerosis: evidence for the implication of the p53 signaling pathway. Neurobiol. Dis., 7, 406–15CrossRefGoogle Scholar
Deak, J. C., Cross, J. V., Lewis, M.et al. (1998). Fas-induced proteolytic activation and intracellular redistribution of the stress-signaling kinase MEKK1. Proc. Natl. Acad. Sci. USA, 95, 5595–600CrossRefGoogle ScholarPubMed
Deckwerth, T. L., Elliott, J. L., Knudson, C. M., Johnson, E. M., Snider, W. D., and Korsmeyer, S. J. (1996). Bax is required for neuronal death after trophic factor deprivation and during development. Neuron, 17, 401–11CrossRefGoogle ScholarPubMed
DeKosky, S. T. and Scheff, S. W. (1990). Synapse loss in frontal cortex biopsies in Alzheimer's disease: correlation with cognitive severity. Ann. Neurol., 27, 457–64CrossRefGoogle ScholarPubMed
Monte, S., Sohn, Y. K., and Wands, J. R. (1997). Correlates of p53- and Fas (CD95)-mediated apoptosis in Alzheimer's disease. J. Neurol. Sci., 152, 73–83CrossRefGoogle ScholarPubMed
Monte, S. M., Sohn, Y. K., Ganju, N., and Wands, J. R. (1998). p53- and CD95-associated apoptosis in neurodegenerative diseases. Lab. Invest., 78, 401–11Google ScholarPubMed
del Peso, L., Gonzlez-Garcia, M., Page, C., Herrera, R., and Nuez, G. (1997). Interleukin-3-induced phosphorylation of bad through the protein kinase Akt. Science, 278, 687–9CrossRefGoogle ScholarPubMed
Deng, H.-X., Hentati, A., Tainer, J. A.et al. (1993). Amyotrophic lateral sclerosis and structural defects in Cu, Zn superoxide dismutase. Science, 261, 1047–51CrossRefGoogle ScholarPubMed
Desagher, S., Osen-Sand, A., Nichols, A.et al. (1999). Bid-induced conformational change of bax is responsible for mitochondrial cytochrome c release during apoptosis. J. Cell Biol., 144, 891–901CrossRefGoogle ScholarPubMed
Dessi, F., Charriaut-Marlangue, C., Khrestchatisky, M., and Ben-Ari, Y. (1993). Glutamate-induced neuronal death is not a programmed cell death in cerebellar culture. J. Neurochem., 60, 1953–5CrossRefGoogle Scholar
Deveraux, Q. L., Roy, N., Stennicke, H. R.et al. (1998). IAPs block apoptotic events induced by caspase-8 and cytochrome c by direct inhibition of distinct caspases. EMBO J., 17, 2215–23CrossRefGoogle ScholarPubMed
Di Cristofano, A., Pesce, B., Cordon-Cardo, C., and Pandolfi, P. P. (1998). Pten is essential for embryonic development and tumor suppression. Nat. Genet., 19, 348–55CrossRefGoogle Scholar
Du, C., Fang, M., Li, Y., Li, L., and Wang, X. (2000). Smac, a mitochondrial protein that promotes cytochrome c-dependent caspase activation by eliminating IAP inhibition. Cell, 102, 33–42CrossRefGoogle ScholarPubMed
Dubois-Dauphin, M., Frankowski, H., Tsujimoto, Y., Huarte, J., and Martinou, J.-C. (1994). Neonatal motoneurons overexpressing the bcl-2 protooncogene in transgenic mice are protected from axotomy-induced cell death. Proc. Natl. Acad. Sci. USA, 91, 3309–13CrossRefGoogle ScholarPubMed
Duker, N. J., Sperling, J., Soprano, K. J., Druin, D. P., Davis, A., and Ashworth, R. (2001). β-amyloid protein induces the formation of purine dimers in cellular DNA. J. Cell Biochem., 81, 393–4003.0.CO;2-5>CrossRefGoogle ScholarPubMed
English, J., Pearson, G., Wilsbacher, J.et al. (1999). New insights into the control of MAP kinase pathways. Exp. Cell Res., 253, 255–70CrossRefGoogle ScholarPubMed
Ethell, D. W., Kinloch, R., and Green, D. R. (2002). Metalloproteinase shedding of Fas ligand regulates beta-amyloid neurotoxicity. Curr. Biol., 12, 1595–600CrossRefGoogle ScholarPubMed
Evans, D. A., Funkenstein, H. H., Albert, M. S.et al. (1989). Prevalence of Alzheimer's disease in a community population of older persons: higher than previously reported. JAMA, 262, 2551–6CrossRefGoogle Scholar
Fan, Z., Beresford, P. J., Oh, D. Y., Zhang, D., and Lieberman, J. (2003). Tumor suppressor NM23-H1 is a granzyme A-activated DNase during TL-mediated apoptosis, and the nucleosome assembly protein SET is its inhibitor. Cell, 112, 659–72CrossRefGoogle Scholar
Farlie, P. G., Dringen, R., Rees, S. M., Kannourakis, G., and Bernard, O. (1995). Bcl-2 transgene expression can protect neurons against developmental and induced cell death. Proc. Natl. Acad. Sci. USA, 92, 4397–401CrossRefGoogle ScholarPubMed
Fath, T., Eidenmuller, J., and Brandt, R. (2002). Tau-mediated cytotoxicity in a pseudohyperphosphorylation model of Alzheimer's disease. J. Neurosci., 22, 9733–41CrossRefGoogle Scholar
Fernandes, R. S. and Cotter, T. G. (1994). Apoptosis or necrosis: intracellular levels of glutathione influence mode of cell death. Biochem. Pharmacol., 48, 675–81CrossRefGoogle ScholarPubMed
Ferrante, R. J., Browne, S. E., Shinobu, L. A.et al. (1997). Evidence of increased oxidative damage in both sporadic and familial amyotrophic lateral sclerosis. J. Neurochem., 69, 2064–74CrossRefGoogle ScholarPubMed
Ferrer, I., Puig, B., Krupinski, J., Carmona, M., and Blanco, R. (2001). Fas and Fas ligand expression in Alzheimer's disease. Acta. Neuropathol., 102, 121–31Google ScholarPubMed
Fitzmaurice, P. S., Shaw, I. C., Kleiner, H. E.et al. (1996). Evidence for DNA damage in amyotrophic lateral sclerosis. Muscle. Nerve., 19, 797–8Google ScholarPubMed
Friedlander, R. M., Brown, R. H., Gagliardini, V., Wang, J., and Yuan, J. (1997). Inhibition of ICE slows ALS in mice. Nature, 388, 31CrossRefGoogle ScholarPubMed
Froelich, S., Basun, H., Forsell, C.et al. (1997). Mapping of a disease locus for familial rapidly progressive frontotemporal dementia to chromosome 17q12–21. Am. J. Med. Genet., 74, 380–53.0.CO;2-T>CrossRefGoogle ScholarPubMed
Gamliel, A., Teicher, C., Hartmann, T., Beyreuther, K., and Stein, R. (2003). Overexpression of wild-type presenilin 2 or its familial Alzheimer's disease-associated mutant does not induce or increase susceptibility to apoptosis in different cells. Neuroscience, 117, 119–28CrossRefGoogle Scholar
Gastard, M. C., Troncoso, J. C., and Koliatsos, V. E. (2003). Caspase activation in the limbic cortex of subjects with early Alzheimer's disease. Ann. Neurol., 54, 393–8CrossRefGoogle ScholarPubMed
Gervais, F. G., Singaraja, R., Xanthoudakis, S.et al. (2002). Recruitment and activation of caspase-8 by the Huntingtin-interacting protein Hip-1 and a novel partner Hippi. Nat. Cell Biol., 4, 95–105CrossRefGoogle Scholar
Gervais, F. G., Xu, D., Robertson, G. S.et al. (1999). Involvement of caspases in proteolytic cleavage of Alzheimer's amyloid-beta precursor protein and amyloidogenic A beta peptide formation. Cell, 97, 395–406CrossRefGoogle ScholarPubMed
Giaccia, A. J. and Kastan, M. B. (1998). The complexity of p53 modulation: emerging patterns from divergent signals. Genes Develop., 12, 2973–83CrossRefGoogle ScholarPubMed
Giasson, B. I., Duda, J. E., Murray, I. V. J.et al. (2000). Oxidative damage linked to neurodegeneration by selective alpha-synuclein nitration in synucleinopathy lesions. Science, 290, 985–9CrossRefGoogle ScholarPubMed
Gilman, C. P., Chan, S. L., Guo, Z., Zhu, X., Greig, N., and Mattson, M. P. (2003). p53 is present in synapses where it mediates mitochondrial dysfunction and synaptic degeneration in response to DNA damage, and oxidative and excitotoxic insults. Neuromolecular Med., 3, 159–72CrossRefGoogle ScholarPubMed
Giovanni, A., Keramaris, E., Morris, E. J.et al. (2000). E2F1 mediates death of β-amyloid-treated cortical neurons in a manner independent of p53 and dependent on Bax and caspase 3. J. Biol. Chem., 275, 11553–60CrossRefGoogle Scholar
Glücksmann, A. (1951). Cell deaths in normal vertebrate ontogeny. Biol. Rev., 26, 59–86CrossRefGoogle ScholarPubMed
Goate, A., Chartier-Harlin, M.-C., Mullan, M.et al. (1991). Segregation of a missense mutation in the amyloid precursor protein gene with familial Alzheimer's disease. Nature, 349, 704–6CrossRefGoogle ScholarPubMed
Golden, W. C., Brambrink, A. M., Traystman, R. J., and Martin, L. J. (2001). Failure to sustain recovery of Na, K ATPase function is a possible mechanism for striatal neurodegeneration in hypoxic-ischemic newborn piglets. Mol. Brain Res., 88, 94–102CrossRefGoogle ScholarPubMed
Goldschmidt, R. B. and Steward, O. (1980). Preferential neurotoxicity of colchicine for granule cells of the dentate gyrus of adult rat. Proc. Natl. Acad. Sci. USA, 77, 3047–51CrossRefGoogle ScholarPubMed
Gotz, J., Chen, F., Barmettler, R., and Nitsch, R. M. (2001a). Tau filament formation in transgenic mice expressing P301L tau. J. Biol. Chem., 276, 529–34CrossRefGoogle Scholar
Gotz, J., Chen, F., Dorpe, J., and Nitsch, R. M. (2001b). Formation of neurofibrillary tangles in P301l tau transgenic mice induced by Abeta 42 fibrils. Science, 293, 1491–5CrossRefGoogle Scholar
Gomez-Isla, T., Price, J. L., McKeel, D. W. Jr.., Growdon, J. H., and Hyman, B. T. (1996). Profound loss of layer II entorhinal cortex neurons occurs in very mild Alzheimer's disease. J. Neurosci., 16, 4491–500CrossRefGoogle ScholarPubMed
Guegan, C., Vila, M., Rosoklija, G., Hays, A. P., and Przeborski, S. (2001). Recruitment of the mitochondrial-dependent apoptotic pathway in amyotrophic lateral sclerosis. J. Neurosci., 21, 6569–76CrossRefGoogle ScholarPubMed
Gwag, B. J., Koh, J. Y., DeMaro, J. A., Ying, H. S., Jacquin, M., and Choi, D. W. (1997). Slowly triggered excitotoxicity occurs by necrosis in cortical cultures. Neuroscience, 77, 393–401CrossRefGoogle ScholarPubMed
Hackam, A. S., Yassa, A. S., Singaraja, R.et al. (2000). Huntingtin interacting protein 1 induces apoptosis via a novel caspase-dependent death effector domain. J. Biol. Chem., 275, 41299–308CrossRefGoogle Scholar
Hakem, R., Hakem, A., Duncan, G. S.et al. (1998). Differential requirement for caspase 9 in apoptotic pathways in vivo. Cell, 94, 339–52CrossRefGoogle ScholarPubMed
Haldar, S., Jena, N., and Croce, C. M. (1995). Inactivation of Bcl-2 by phosphorylation. Proc. Natl. Acad. Sci. USA, 92, 4507–11CrossRefGoogle ScholarPubMed
Hardy, J. and Selkoe, D. J. (2002). The amyloid hypothesis of Alzheimer's disease: progress and problems on the road to therapeutics. Science, 297, 353–6CrossRefGoogle ScholarPubMed
Harper, P. S. (1996). Huntington's Disease. London: W. B. Saunders
Hartmann, A., Hunot, S., Michel, P. P.et al. (2000). Caspase-3: a vulnerability factor and final effector in apoptotic death of dopaminergic neurons in Parkinson's disease. Proc. Natl. Acad. Sci., 97, 2875–80CrossRefGoogle ScholarPubMed
Hartmann, A., Michel, P. P., Troadec, J.-D.et al. (2001). Is Bax a mitochondrial mediator in apoptotic death of dopaminergic neurons in Parkinson's disease?J. Neurochem., 76, 1785–93CrossRefGoogle Scholar
Hayward, C., Colville, S., Swingler, R. J., and Brock, D. J. H. (1999). Molecular genetic analysis of the APEX nuclease gene in amyotrophic lateral sclerosis. Neurology, 52, 1899–901CrossRefGoogle ScholarPubMed
He, Y., Lee, T., and Leong, S. K. (2000). 6-hydroxydopamine induced apoptosis of dopaminergic cells in the rat substantia nigra. Brain Res., 858, 163–6CrossRefGoogle ScholarPubMed
Hegde, R., Srinivasula, S. M., Zhang, Z.et al. (2002). Identification of Omi/HtrA2 as a mitochondrial apoptotic serine protease that disrupts inhibitor of apoptosis protein-caspase interaction. J. Biol. Chem., 277, 432–8CrossRefGoogle ScholarPubMed
Hetman, M., Kanning, K., Cavanaugh, J. E., and Xia, Z. (1999). Neuroprotection by brain-derived neurotrophic factor is mediated by extracellular signal-regulated kinase and phosphatidylinositol 3-kinase. J. Biol. Chem., 2744, 22569–80CrossRefGoogle Scholar
Hirsch, T., Marchetti, P., Susin, S. A.et al. (1997). The apoptosis-necrosis paradox. Apoptogenic proteases activated after mitochondrial permeability transition determine the mode of cell death. Oncogene, 15, 1573–81CrossRefGoogle ScholarPubMed
Hodgson, J. G., Agopyan, N., Gutekunst, C. A.et al. (1999). A YAC mouse model for Huntington's disease with full-length mutant huntingtin, cytoplasmic toxicity, and selective striatal neurodegeneration. Neuron, 23, 181–92CrossRefGoogle ScholarPubMed
Holcik, M. (2002). The IAP proteins. Trends Gen., 18, 537–8CrossRefGoogle ScholarPubMed
Holcik, M., Thompson, C. S., Yaraghi, Z., Lefebvre, C. A., MacKenzie, A. E., and Korneluk, R. G. (1999). The hippocampal neurons of neuronal apoptosis inhibitory protein 1 (NAIP1)-deleted mice display increased vulnerability to kainic acid-induced injury. Proc. Natl. Acad. Sci. USA, 97, 2286–90CrossRefGoogle Scholar
Holzer, M., Gartner, U., Klinz, F. J., Narz, F., Heumann, R., and Arendt, T. (2001). Activation of mitogen-activated protein kinase cascade and phosphorylation of cytoskeletal proteins after neurone-specific activation of p21ras. I. Mitogen-activated protein kinase cascade. Neuroscience, 105, 1031–40CrossRefGoogle ScholarPubMed
Hu, Y., Benedict, M. A., Wu, D., Inohara, N., and Núñez, G. (1998). Bcl-xL interacts with Apaf-1 and inhibits Apaf-1-dependent caspase-9 activation. Proc. Natl. Acad. Sci. USA, 95, 4386–91CrossRefGoogle ScholarPubMed
The Huntington's Disease Collaborative Research Group (1993). A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. Cell, 72, 971–83CrossRef
Ikonomidou, C., Qin, Y. Q., Labruyere, J., and Olney, J. W. (1996). Motor neuron degeneration induced by excitotoxin agonists has features in common with those seen in the SOD-1 transgenic mouse model of amyotrophic lateral sclerosis. J. Neuropathol. Exp. Neurol., 55, 211–44CrossRefGoogle ScholarPubMed
Inbal, B., Bialik, S., Sabanay, I., Shani, G., and Kimchi, A. (2002). DAP kinase and DRP-1 mediate membrane blebbing and the formation of autophagic vesicles during programmed cell death. J. Cell Biol., 157, 455–68CrossRefGoogle ScholarPubMed
Irizarry, M. C., McNamara, M., Fedorchak, K., Hsiao, K., and Hyman, B. T. (1997). APPSw transgenic mice develop age-related A beta deposits and neuropil abnormalities, but no neuronal loss in CA1. J. Neuropathol. Exp. Neurol., 56, 965–73CrossRefGoogle ScholarPubMed
Ishikawa, Y. and Kitamura, M. (1999). Dual potential of extracellular signal-regulated kinase for the control of cell survival. Biochem. Biophys. Res. Comm., 2644, 696–701CrossRefGoogle Scholar
Iwata, A., Maruyama, M., Kanazawa, I., and Nukina, N. (2001). α-synuclein affects the MAPK pathway and accelerates cell death. J. Biol. Chem., 276, 45320–9CrossRefGoogle ScholarPubMed
Jellinger, K. A. (1999). Is there apoptosis in Lewy body disease?Acta. Neuropathol., 97, 413–15CrossRefGoogle ScholarPubMed
Johnson, M. D., Kinoshita, Y., Xiang, H., Ghatan, S., and Morrison, R. S. (1999). Contribution of p53-dependent caspase activation to neuronal cell death declines with neuronal maturation. J. Neurosci., 19, 2996–3006CrossRefGoogle ScholarPubMed
Jones, J. M., Datta, P., Srinivasula, S. M.et al. (2003). Loss of Omi mitochondrial protease activity causes the neuromuscular disorder of mnd2 mutant mice. Nature, 425, 721–7CrossRefGoogle ScholarPubMed
Jost, C. A., Marin, M. C., and Kaelin, W. G. (1997). p73 is a human p53-related protein that can induce apoptosis. Nature, 389, 191–4CrossRefGoogle ScholarPubMed
Kalaria, R. N. (2003). Dementia comes of age in the developing world. Lancet, 361, 888–9CrossRefGoogle ScholarPubMed
Kaplan, D. R. and Miller, F. D. (1997). Signal transduction by the neurotrophin receptors. Curr. Opin. Cell Biol., 9, 213–21CrossRefGoogle ScholarPubMed
Katzman, R. (1993). Education and the prevalence of dementia and Alzheimer's disease. Neurology, 43, 13–20CrossRefGoogle ScholarPubMed
Keilin, D. (1930). Cytochrome and intracellular oxidase. Proc. R. Soc. Lond., 106, 418–44CrossRefGoogle Scholar
Kemp, J. A. and McKernan, R. M. (2002). NMDA receptor pathways as drug targets. Nat. Neurosci. (suppl), 5, 1039–42CrossRefGoogle ScholarPubMed
Kerr, J. F. R., Wyllie, A. H., and Currie, A. R. (1972). Apoptosis: a basic biological phenomenon with wide-ranging implications in tissue kinetics. Br. J. Cancer, 26, 239–57CrossRefGoogle ScholarPubMed
Kim, J. A., Mitsukawa, K., Yamada, M. K., Nishiyama, N., Matsuki, N., and Ikegaya, Y. (2002). Cytoskeleton disruption causes apoptotic degeneration of dentate granule cells in hippocampal slice cultures. Neuropharmacology, 42, 1109–18CrossRefGoogle ScholarPubMed
Kim, M., Lee, H. S., LaForet, G.et al. (1999). Mutant huntingtin expression in clonal striatal cells: dissociation of inclusion formation and neuronal survival by caspase inhibition. J. Neurosci., 19, 964–73CrossRefGoogle ScholarPubMed
Kim, T.-W., Pettingell, W. H., Jung, Y.-K., Kovacs, D. M., and Tazi, R. E. (1997). Alternative cleavage of Alzheimer-associated presenilins during apoptosis by a caspase-3 family protein. Science, 277, 373–6CrossRefGoogle Scholar
Kisby, G. E., Miline, J., and Sweatt, C. (1997). Evidence of reduced DNA repair in amyotrophic lateral sclerosis brain tissue. NeuroReport, 8, 1337–40CrossRefGoogle ScholarPubMed
Kitamura, Y., Shimohama, S., Kamoshima, W., Matsuoka, Y., Nomura, Y., and Taniguchi, T. (1997). Changes of p53 in the brains of patients with Alzheimer's disease. Biochem. Biophys. Res. Comm., 232, 418–21CrossRefGoogle ScholarPubMed
Kitamura, Y., Shimohama, S., Kamoshima, W.et al. (1998). Alteration of proteins regulating apoptosis, Bcl-2, Bcl-x, Bax, Bak, Bad, ICH-1 and CPP32, in Alzheimer's disease. Brain Res., 780, 260–9CrossRefGoogle ScholarPubMed
Klein, J. A., Longo-Guess, C. M., Rossmann, M. P.et al. (2002). The harlequin mouse mutation downregulates apoptosis-inducing factor. Nature, 419, 367–74CrossRefGoogle ScholarPubMed
Klionsky, D. J. and Emr, S. D. (2000). Autophagy as a regulated pathway of cellular degradation. Science, 290, 1717–21CrossRefGoogle ScholarPubMed
Kluck, R. M., Bossy-Wetzel, E., Green, D. R., and Newmeyer, D. D. (1997). The release of cytochrome c from mitochondria: a primary site for bcl-2 regulation of apoptosis. Science, 275, 1132–6CrossRefGoogle ScholarPubMed
Kosik, K. S., Joachim, C. L., and Selkoe, D. J. (1986). Microtubule-associated protein tau is a major antigenic component of paired helical filaments in Alzheimer's disease. Proc. Natl. Acad. Sci. USA, 83, 4044–8CrossRefGoogle Scholar
Kostic, V., Jackson-Lewis, V., Bilbao, F., Dubois-Dauphin, M., and Przedborski, S. (1997). Bcl-2: prolonging life in a transgenic mouse model of familial amyotrophic lateral sclerosis, Science, 277, 559–62CrossRefGoogle Scholar
Kuida, K., Haydar, T. F., Kuan, C.-Y.et al. (1998). Reduced apoptosis and cytochrome c-mediated caspase activation in mice lacking caspase-9. Cell, 94, 325–37CrossRefGoogle ScholarPubMed
Kuntz, C., Kinoshita, Y., Beal, M. F., Donehower, L. A., and Morrison, R. S. (2000). Absence of p53: no effect in a transgenic mouse model of familial amyotrophic lateral sclerosis. Exp. Neurol., 165, 184–90CrossRefGoogle Scholar
Kuperstein, F. and Yavin, E. (2002). ERK activation and nuclear translocation in amyloid-beta peptide- and iron-stressed neuronal cell cultures. Eur. J. Neurosci., 16, 44–54CrossRefGoogle ScholarPubMed
Kure, S., Tominaga, T., Yoshimoto, T., Tada, K., and Narisawa, K. (1991). Glutamate triggers internucleosomal DNA cleavage in neuronal cells. Biochem. Biophys. Res. Commun., 179, 39–45CrossRefGoogle ScholarPubMed
Kwon, C. H., Zhu, X., Zhang, J.et al. (2001). Pten regulates neuronal soma size: a mouse model of Lhermitte–Duclos disease. Nat. Genet., 29, 410–11CrossRefGoogle ScholarPubMed
LaCasse, E. C., Baird, S., Korneluk, R. G., and MacKenzie, A. E. (1998). The inhibitors of apoptosis (IAPs) and their emerging role in cancer. Oncogene, 17, 3247–59CrossRefGoogle Scholar
Lachyankar, M. B., Condon, P. J., Quesenberry, P. J.et al. (2000). A role for nuclear PTEN in neuronal differentiation. J. Neurosci., 20, 1404–13CrossRefGoogle ScholarPubMed
LaFerla, F. M., Hall, C. K., Ngo, L., and Jay, G. (1996). Extracellular deposition of β-amyloid upon p53-dependent neuronal cell death in transgenic mice. J. Clin. Invest., 98, 1626–32CrossRefGoogle ScholarPubMed
Lambrechts, D., Storkebaum, E., Morimoto, M.et al. (2003). VEGF is a modifier of amyotrophic lateral sclerosis in mice and humans and protects motoneurons against ischemic death. Nat. Gen., 34, 383–94CrossRefGoogle ScholarPubMed
Lashley, K. S. (1941). Thalamo-cortical connections of the rat's brain. J. Comp. Neurol., 75, 67–121CrossRefGoogle Scholar
LeBlanc, A., Liu, H., Goodyer, C., Bergeron, C., and Hammond, J. (1999). Caspase-6 role in apoptosis of human neurons, amyloidogenesis, and Alzheimer's disease. J. Biol. Chem., 274, 23426–36CrossRefGoogle ScholarPubMed
Leist, M., Single, B., Castoldi, A. F., Kühnle, S., and Nicotera, P. (1997). Intracellular adenosine triphosphate (ATP) concentration: a switch in the decision between apoptosis and necrosis. J. Exp. Med., 185, 1481–6CrossRefGoogle ScholarPubMed
Lennon, S. V., Martin, S. J., and Cotter, T. G. (1991). Dose-dependent induction of apoptosis in human tumour cell lines by widely diverging stimuli. Cell Prolif., 24, 203–14CrossRefGoogle ScholarPubMed
Leppin, C., Finiels-Marlier, F., Crawley, J. N., Montpied, P., and Paul, S. M. (1992). Failure of a protein synthesis inhibitor to modify glutamate receptor-mediated neurotoxicity in vivo. Brain Res., 581, 168–70CrossRefGoogle ScholarPubMed
Lesuisse, C. and Martin, L. J. (2002). Immature and mature neurons engage different apoptotic mechanisms involving caspase-3 and the mitogen-activated protein kinase pathway. J. Cereb. Blood Flow Metabol., 22, 935–50CrossRefGoogle ScholarPubMed
Letai, A., Bassik, M. C., Walensky, L. D., Sorcinelli, M. D., Weiler, S., and Korsmeyer, S. J. (2002). Distinct BH3 domains either sensitize or activate mitochondrial apoptosis, serving as prototype cancer therapeutics. Cancer Cell, 2, 183–92CrossRefGoogle ScholarPubMed
Levrero, M., Laurenzi, V., Costanzo, A.et al. (2000). The p53/p63/p73 family of transcription factors: overlapping and distinct functions. J. Cell Sci., 113, 1661–70Google ScholarPubMed
Lewis, J., Dickson, D. W., Lin, W. L.et al. (2001). Enhanced neurofibrillary degeneration in transgenic mice expressing mutant tau and APP. Science, 293, 1487–91CrossRefGoogle ScholarPubMed
Li, H., Zhu, H., Xu, C.-J., and Yuan, J. (1998). Cleavage of Bid by caspase 8 mediates the mitochondrial damage in the Fas pathway of apoptosis. Cell, 94, 491–501CrossRefGoogle Scholar
Li, M., Ona, V. O., Guegan, C.et al. (2000). Functional role of caspase-1 and caspase-3 in an ALS transgenic mouse model. Science, 288, 283–4CrossRefGoogle Scholar
Li, P., Nijhawan, D., Budihardjo, I.et al. (1997). Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an apoptotic protease cascade. Cell, 91, 479–89CrossRefGoogle ScholarPubMed
Liange, X. H., Kleeman, L. K., Jiang, H. H.et al. (1998). Protection against fatal Sindbis virus encephalitis by beclin, a novel Bcl-2-interacting protein. J. Virol., 72, 8586–96Google Scholar
Lieberman, A. R. (1971). The axon reaction: a review of the principal features of perikaryal responses to axon injury. Int. Rev. Neurobiol., 14, 49–124CrossRefGoogle ScholarPubMed
Lipton, S. A. and Rosenberg, P. A. (1994). Excitatory amino acids as a final common pathway for neurologic disorders. N. Engl. J. Med., 330, 613–22Google ScholarPubMed
Liston, P., Roy, N., Tamai, K.et al. (1996). Suppression of apoptosis in mammalian cells by NAIP and a related family of IAP genes. Nature, 379, 349–53CrossRefGoogle Scholar
Lithgow, T., Driel, R., Bertram, J. F., and Strasser, A. (1994). The protein product of the oncogene bcl-2 is a component of the nuclear envelope, the endoplasmic reticulum, and the outer mitochondrial membrane. Cell Growth Differ., 5, 411–17Google ScholarPubMed
Liu, X., Kim, C. N., Yang, J., Jemmerson, R., and Wang, X. (1996). Induction of apoptotic program in cell-free extracts: requirement for dATP and cytochrome c. Cell, 86, 147–57CrossRefGoogle ScholarPubMed
Liu, X., Zou, H., Slaughter, C., and Wang, X. (1997). DFF, a heterodimeric protein that functions downstream of caspase-3 to trigger DNA fragmentation during apoptosis. Cell, 89, 175–84CrossRefGoogle ScholarPubMed
Liu, Z., Gastard, M., Verina, T., Bora, S., Mouton, P. R., and Koliatsos, V. E. (2001). Estrogens modulate experimentally induced apoptosis of granule cells in the adult hippocampus. J. Comp. Neurol., 441, 1–8CrossRefGoogle ScholarPubMed
Lockshin, R. A. and Williams, C. M. (1964). Programmed cell death: II. Endocrine potentiation of the breakdown of the intersegmental muscles of silkmoths. J. Insect. Physiol., 10, 643–9CrossRefGoogle Scholar
Lockshin, R. A. and Zakeri, Z. (2001). Programmed cell death and apoptosis: origins of the theory. Nat. Rev. Mol. Cell Biol., 2, 545–50CrossRefGoogle Scholar
Lockshin, R. A. and Zakeri, Z. (2002). Caspase-independent cell deaths. Curr. Opin. Cell Biol., 14, 727–33CrossRefGoogle ScholarPubMed
Lok, J. and Martin, L. J. (2002). Rapid subcellular redistribution of Bax precedes caspase-3 and endonuclease activation during excitotoxic neuronal apoptosis in rat brain. J. Neurotrauma, 19, 815–28CrossRefGoogle ScholarPubMed
Loo, D. T., Copani, A., Pike, C. J., Whittemore, E. R., Walencewicz, A. J., and Cotman, C. W. (1993). Apoptosis is induced by β-amyloid in cultured central nervous system neurons. Proc. Natl. Acad. Sci. USA, 90, 7951–5CrossRefGoogle ScholarPubMed
Lucas, D. R. and Newhouse, J. P. (1957). The toxic effect of sodium L-glutamate on the inner layers of the retina. Arch. Ophthal., 58, 193–201CrossRefGoogle ScholarPubMed
Lucassen, P. J., Chung, W. C. J., Kamphorst, W., and Swaab, D. F. (1997). DNA damage distribution in the human brain as shown by in situ end labeling: area-specific differences in aging and Alzheimer's disease in the absence of apoptotic morphology. J. Neuropathol. Exp. Neurol., 56, 887–900CrossRefGoogle ScholarPubMed
Majno, G. and Joris, I. (1995). Apoptosis, oncosis, and necrosis. An overview of cell death. Am. J. Pathol., 146, 3–15Google ScholarPubMed
Maki, C. G., Huibregtse, J. M., and Howley, P. M. (1996). In vivo ubiquitination and proteasome-mediated degradation of p53. Cancer Res., 56, 2649–54Google ScholarPubMed
Martin, L. J. (1999). Neuronal death in amyotrophic lateral sclerosis is apoptosis: possible contribution of a programmed cell death mechanism. J. Neuropathol. Exp. Neurol., 58, 459–71CrossRefGoogle ScholarPubMed
Martin, L. J. (2000). p53 is abnormally elevated and active in the CNS of patients with amyotrophic lateral sclerosis. Neurobiol. Dis., 7, 613–22CrossRefGoogle ScholarPubMed
Martin, L. J. (2001). Neuronal cell death in nervous system development, disease, and injury. Int. J. Mol. Med., 7, 455–78Google ScholarPubMed
Martin, L. J. and Liu, Z. (2002). Injury-induced spinal motor neuron apoptosis is preceded by DNA single-strand breaks and is p53- and Bax-dependent. J. Neurobiol., 50, 181–97CrossRefGoogle ScholarPubMed
Martin, L. J., Al-Abdulla, N. A., Brambrink, A. M., Kirsch, J. R., Sieber, F. E. and Portera-Cailliau, C. (1998). Neurodegeneration in excitotoxicity, global cerebral ischemia, and target deprivation: a perspective on the contributions of apoptosis and necrosis. Brain Res. Bull., 46, 281–309CrossRefGoogle ScholarPubMed
Martin, L. J., Kaiser, A., and Price, A. C. (1999). Motor neuron degeneration after sciatic nerve avulsion in adult rat evolves with oxidative stress and is apoptosis. J. Neurobiol., 40, 185–2013.0.CO;2-#>CrossRefGoogle ScholarPubMed
Martin, L. J., Kaiser, A., Yu, J. W., Natale, J. E., and Al-Abdulla, N. A. (2001). Injury-induced apoptosis of neurons in adult brain is mediated by p53-dependent and p53-independent pathways and requires Bax. J. Comp. Neurol., 433, 299–311CrossRefGoogle ScholarPubMed
Martin, L. J., Pardo, C. A., Cork, L. C., and Price, D. L. (1994). Synaptic pathology and glial responses to neuronal injury precede the formation of senile plaques and amyloid deposits in the aging cerebral cortex. Am. J. Pathol., 145, 1358–81Google ScholarPubMed
Martin, L. J., Price, A. C., McClendon, K. B.et al. (2003). Early events of target deprivation/axotomy-induced neuronal apoptosis in vivo: oxidative stress, DNA damage, p53 phosphorylation and subcellular redistribution of death proteins. J. Neurochem., 85, 234–47CrossRefGoogle ScholarPubMed
Martin, L. J., Sieber, F. E., and Traystman, R. J. (2000a). Apoptosis and necrosis occur in separate neuronal populations in hippocampus and cerebellum after ischemia and are associated with alterations in metabotropic glutamate receptor signaling pathways. J. Cereb. Blood Flow Metab., 20, 153–67CrossRefGoogle Scholar
Martin, L. J., Price, A. C., Kaiser, A., Shaikh, A. Y., and Liu, Z. (2000b). Mechanisms for neuronal degeneration in amyotrophic lateral sclerosis and in models of motor neuron death. Int. J. Mol. Med., 5, 3–13Google Scholar
Martinou, J.-C., Dubois-Dauphin, M., Staple, J. K.et al. (1994). Overexpression of bcl-2 in transgenic mice protects neurons from naturally occurring cell death and experimental ischemia. Neuron, 13, 1017–30CrossRefGoogle ScholarPubMed
Marzo, I., Brenner, C., Zamzami, N.et al. (1998). Bax and adenine nucleotide translocator cooperate in the mitochondrial control of apoptosis. Science, 281, 2027–31CrossRefGoogle Scholar
Mate, M. J., Ortiz-Lombardia, M., Boitel, B.et al. (2002). The crystal structure of the mouse apoptosis-inducing factor AIF. Nature Struct. Biol., 9, 442–6CrossRefGoogle ScholarPubMed
Mattson, M. P., Cheng, B., Davis, D., Bryant, K., Lieberburg, I., and Rydel, R. E. (1992). β-amyloid peptides destabilize calcium homeostasis and render human cortical neurons vulnerable to excitoxicity. J. Neurosci., 12, 376–89CrossRefGoogle Scholar
McKhann, G., Drachman, D., Folstein, M., Katzman, R., Price, D., and Stadlan, E. M. (1984). Clinical diagnosis of Alzheimer's disease: report of the NINCDS-ADRDA work group under the auspices of the Department of Health and Human Services task force on Alzheimer's disease. Neurology, 34, 939–44CrossRefGoogle ScholarPubMed
McPhail, L. T., Vanderluit, J. L., McBride, C. B.et al. (2003). Endogenous expression of inhibitor of apoptosis proteins in facial motoneurons of neonatal and adult rats following axotomy. Neuroscience, 117, 567–75CrossRefGoogle ScholarPubMed
Merry, D. E. and Korsmeyer, S. J. (1997). Bcl-2 gene family in the nervous system. Ann. Rev. Neurosci., 20, 245–67CrossRefGoogle Scholar
Michaelidis, T. M., Sendtner, M., Cooper, J. D.et al. (1996). Inactivation of bcl-2 results in progressive degeneration of motoneurons, sympathetic and sensory neurons during early postnatal development. Neuron, 17, 75–89CrossRefGoogle ScholarPubMed
Migheli, A., Atzori, C., Piva, R.et al. (1999). Lack of apoptosis in mice with ALS. Nature Med., 5, 966–7CrossRefGoogle ScholarPubMed
Miller, T. M., Moulder, K. L., Knudson, C. M.et al. (1997). Bax deletion further orders the cell death pathway in cerebellar granule cells and suggests a caspase-independent pathway to cell death. J. Cell Biol., 139, 205–17CrossRefGoogle ScholarPubMed
Mizushima, N., Ohsumi, Y., and Yoshimori, T. (2002). Autophagosome formation in mammalian cells. Cell Struct. Funct., 27, 421–9CrossRefGoogle ScholarPubMed
Morishima, Y., Gotoh, Y., Zieg, J.et al. (2001). β-amyloid induces neuronal apoptosis via a mechanism that involves the c-Jun N-terminal kinase pathway and the induction of Fas ligand. J. Neurosci., 21, 7551–60CrossRefGoogle Scholar
Mouton, P. R., Martin, L. J., Calhoun, M. E., Dal Forno, G., and Price, D. L. (1998). Cognitive decline strongly correlates with cortical atrophy in Alzheimer's disease. Neurobiol. Aging, 19, 371–7CrossRefGoogle Scholar
Muchmore, S. W., Sattler, M., Liang, H.et al. (1999). X-ray and NMR structure of human Bcl-xL, an inhibitor of programmed cell death. Nature, 381, 335–41CrossRefGoogle Scholar
Mund, T., Gewies, A., Schoenfield, N., Bauer, M. K. A., and Grimm, S. (2003). Spike, a novel BH3-only protein, regulates apoptosis at the endoplasmic reticulum. FASEB J. (February 19). 10.1096/fj.02-0657fjeCrossRef
Nagata, S. (1999). Fas ligand-induced apoptosis. Annu. Rev. Genet., 33, 29–55CrossRefGoogle ScholarPubMed
Nakagawa, T., Zhu, H., Morishima, N.et al. (2000). Caspase-12 mediates endoplasmic reticulum-specific apoptosis and cytotoxicity by amyloid-beta. Nature, 403, 98–103CrossRefGoogle ScholarPubMed
Nakajima, W., Ishida, A., Lange, M. S.et al. (2000). Apoptosis has a prolonged role in the neurodegeneration after hypoxic ischemia in the newborn rat. J. Neurosci., 20, 7994–8004CrossRefGoogle Scholar
Naruse, S., Igarashi, S., Kobayashi, H.et al. (1991). Mis-sense mutation Val→Ile in exon 17 of amyloid precursor protein gene in Japanese familial Alzheimer's disease. Lancet., 337, 978–9CrossRefGoogle ScholarPubMed
Natale, J. E., Cheng, Y., and Martin, L. J. (2002). Thalamic neuron apoptosis emerges rapidly after cortical damage in immature mice. Neuroscience, 112, 665–76CrossRefGoogle ScholarPubMed
Nechushtan, A., Smith, C. L., Lamensdorf, I., Yoon, S.-H., and Youle, R. J. (2001). Bax and Bak coalesce into novel mitochondria-associated clusters during apoptosis. J. Cell Biol., 153, 1265–76CrossRefGoogle ScholarPubMed
Nicholson, S. J., Witherden, A. S., Hafezparast, M., Martin, J. E., and Fisher, E. M. C. (2000). Mice, the motor system, and human motor neuron pathology. Mamm. Genome, 11, 1041–52CrossRefGoogle ScholarPubMed
Northington, F. J., Ferriero, D. M., Graham, E. M., Traystman, R. J., and Martin, L. J. (2001). Early neurodegeneration after hypoxia-ischemia in neonatal rat is necrosis while delayed neuronal death is apoptosis. Neurobiol. Dis., 8, 207–19CrossRefGoogle ScholarPubMed
O'Connor, T. M. and Wyttenbach, C. R. (1974). Cell death in the embryonic chick spinal cord. J. Cell Biol., 60, 448–59CrossRefGoogle ScholarPubMed
Oddo, S., Caccamo, A., Shepherd, J. D.et al. (2003). Triple-transgenic model of Alzheimer's disease with plaques and tangles: intracellular Aβ and synaptic dysfunction. Neuron, 39, 409–21CrossRefGoogle ScholarPubMed
Offen, D., Beart, P. M., Cheung, N. S.et al. (1998). Transgenic mice expressing human Bcl-2 in their neurons are resistant to 6-hydroxydopamine and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine neurotoxicity. Proc. Natl. Acad. Sci., 95, 5789–94CrossRefGoogle ScholarPubMed
Ogier-Denis, E. and Codogno, P. (2003). Autophagy: a barrier or an adaptive response to cancer. Biochim. Biophys. Acta., 1603, 113–28Google ScholarPubMed
Olanow, C. W. and Tatton, W. G. (1999). Etiology and pathogenesis of Parkinson's disease. Ann. Rev. Neurosci., 22, 123–44CrossRefGoogle ScholarPubMed
Olkowski, Z. L. (1998). Mutant AP endonuclease in patients with amyotrophic lateral sclerosis. NeuroReport, 9, 239–42CrossRefGoogle ScholarPubMed
Olney, J. W. (1971). Glutamate-induced neuronal necrosis in the infant mouse hypothalamus. An electron microscopic study. J. Neuropathol. Exp. Neurol., 30, 75–90CrossRefGoogle ScholarPubMed
Olshansky, S. J., Carnes, B. A., and Cassel, C. K. (1993). The aging of the human species. Sci. Am., 268, 46–52CrossRefGoogle ScholarPubMed
Orrenius, S., Zhivotovsky, B., and Nicotera, P. (2003). Regulation of cell death: the calcium-apoptosis link. Nat. Rev., 4, 552–65CrossRefGoogle ScholarPubMed
Paradis, E., Douillard, H., Koutroumanis, M., Goodyer, C., and LeBlanc, A. (1997). Amyloid β peptide of Alzheimer's disease downregulates Bcl-2 and upregulates Bax expression in human neurons. J. Neurosci., 16, 7533–9CrossRefGoogle Scholar
Passer, B. J., Pellegrini, L., Vito, P., Ganjei, J. K., and D'Adamio, L. (1999). Interaction of Alzheimer's presenilin-1 and presenilin-2 with Bcl-X(L). A potential role in modulating the threshold of cell death. J. Biol. Chem., 274, 24007–13CrossRefGoogle ScholarPubMed
Patrick, G. N., Zukerberg, L., Nikolic, M., Monte, S., Dikkes, P., and Tsai, L. H. (1999). Conversion of p35 to p25 deregulates Cdk5 activity and promotes neurodegeneration. Nature, 402, 615–22CrossRefGoogle ScholarPubMed
Pelvig, D. P., Pakkenberg, H., Regeur, L., Oster, S., and Pakkenberg, B. (2003). Neocortical glial cell numbers in Alzheimer's disease. A stereological study. Dement. Geriatr. Cogn. Disord., 16, 212–19CrossRefGoogle ScholarPubMed
Perrelet, D., Ferri, A., Liston, P., Muzzini, P., Korneluk, R. G., and Kato, A. C. (2002). IAPs are essential for GDNF-mediated neuroprotective effects in injured motor neuron in vivo. Nat. Cell Biol., 4, 175–9CrossRefGoogle ScholarPubMed
Petersen, A., Larsen, K. E., Behr, G. G.et al. (2001). Expanded CAG repeats in exon 1 of the Huntington's disease gene stimulate dopamine-mediated striatal neuron autophagy and degeneration. Hum. Mol. Genet., 10, 1243–54CrossRefGoogle ScholarPubMed
Pilar, G. and Landmesser, L. (1976). Ultrastructural differences during embryonic cell death in normal and peripherally deprived ciliary ganglia. J. Cell Biol., 68, 339–56CrossRefGoogle ScholarPubMed
Pompl, P. N., Yemul, S., Xiang, Z.et al. (2003). Caspase gene expression in the brain as a function of the clinical progression of Alzheimer's disease. Arch. Neurol., 60, 369–76CrossRefGoogle Scholar
Poorkaj, P., Bird, T. D., Wijsman, E.et al. (1998). Tau is a candidate gene for chromosome 17 frontotemporal dementia. Ann. Neurol., 43, 815–25CrossRefGoogle ScholarPubMed
Portera-Cailliau, C., Price, D. L., and Martin, L. J. (1997a). Excitotoxic neuronal death in the immature brain is an apoptosis-necrosis morphological continuum. J. Comp. Neurol., 378, 70–87Google Scholar
Portera-Cailliau, C., Price, D. L., and Martin, L. J. (1997b). Non-NMDA and NMDA receptor-mediated excitotoxic neuronal deaths in adult brain are morphologically distinct: further evidence for an apoptosis-necrosis continuum. J. Comp. Neurol., 378, 88–1043.0.CO;2-G>CrossRefGoogle Scholar
Pozniak, C. D., Radinovic, S., Yang, A., McKeon, F., Kaplan, D. R., and Miller, F. D. (2000). An anti-apoptotic role for the p53 family member, p73, during developmental neuron death. Science, 289, 304–6CrossRefGoogle ScholarPubMed
Prestige, M. (1970). Differentiation, degeneration, and the role of the periphery: quantitative considerations. In The Neurosciences: Second Study Program, ed. F. O. Schmitt. New York: Rockefeller University Press, pp. 73–82
Proskuryakov, S. Y., Konoplyannikov, A. G., and Gabai, V. L. (2003). Necrosis: a specific form of programmed cell death. Exp. Cell Res., 283, 1–16CrossRefGoogle ScholarPubMed
Prudlo, J., Koenig, J., Graser, J.et al. (2000). Motor neuron cell death in a mouse model of FALS is not mediated by the p53 cell survival regulator. Brain Res., 897, 183–7CrossRefGoogle Scholar
Putcha, G. V., Deshmukh, M., and Johnson, E. M. Jr. (1999). Bax translocation is a critical event in neuronal apoptosis: regulation by neuroprotectants, Bcl-2, and caspases. J. Neurosci., 19, 7476–85CrossRefGoogle ScholarPubMed
Raffray, M. and Cohen, G. M. (1997). Apoptosis and necrosis in toxicology: a continuum or distinct modes of cell death?Pharmacol. Ther., 75, 153–77CrossRefGoogle ScholarPubMed
Raoul, C., Estvez, A. G., Nishimune, H.et al. (2002). Motoneuron death triggered by a specific pathway downstream of Fas: potentiation by ALS-linked SOD1 mutations. Neuron, 35, 1067–83CrossRefGoogle ScholarPubMed
Rapoport, M., Dawson, H. N., Binder, L. I., Vitek, M. P., and Ferreira, A. (2002). Tau is essential to β-amyloid-induced neurotoxicity. Proc. Natl. Acad. Sci. USA, 99, 6364–9CrossRefGoogle ScholarPubMed
Rathke-Hartlieb, S., Schlomann, U., Heimann, P., Meisler, M. H., Jockusch, H., and Bartsch, J. W. (2002). Progressive loss of striatal neurons causes motor dysfunction in MND2 mutant mice and is not prevented by Bcl-2. Exp. Neurol., 175, 87–97CrossRefGoogle Scholar
Reddy, P. H., Williams, M., Charles, V.et al. (1998). Behavioral abnormalities and selective neuronal loss in HD transgenic mice expressing mutated full-length HD cDNA. Nat. Genet., 20, 198–202CrossRefGoogle Scholar
Robertson, J. D., Enoksson, M., Suomela, M., Zhivotovsky, B., and Orrenius, S. (2002). Caspase-2 acts upstream of mitochondria to promote cytochrome c release during etoposide-induced apoptosis. J. Biol. Chem., 277, 29803–9CrossRefGoogle ScholarPubMed
Robison, S. H., Tandan, R., and Bradley, W. G. (1993). Repair of N-methylpurines in DNA from lymphocytes of patients with amyotrophic lateral sclerosis. J. Neurol. Sci., 115, 201–7CrossRefGoogle ScholarPubMed
Rohn, T. T., Rissman, R. A., Davis, M. C., Kim, Y. E., Cotman, C. W., and Head, E. (2002). Caspase-9 activation and caspase cleavage of tau in the Alzheimer's disease brain. Neurobiol. Dis., 11, 341–54CrossRefGoogle ScholarPubMed
Romanes, G. J. (1946). Motor localization and the effects of nerve injury on the ventral horn cells of the spinal cord. J. Anat., 80, 117–31Google ScholarPubMed
Rosen, D. R., Siddique, T., Patterson, D.et al. (1993). Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature, 362, 59–62CrossRefGoogle ScholarPubMed
Rosenmann, H., Meiner, Z., Kahana, E.et al. (2003). An association study of the codon 72 polymorphism in the pro-apoptotic gene p53 and Alzheimer's disease. Neurosci. Lett., 340, 29–32CrossRefGoogle ScholarPubMed
Roses, A. D. (1996). Apolipoprotein E alleles as risk factors in Alzheimer's disease. Annu. Rev. Med., 47, 387–400CrossRefGoogle ScholarPubMed
Rothstein, J. D., Martin, L. J., and Kuncl, R. W. (1992). Decreased glutamate transport by the brain and spinal cord in amyotrophic lateral sclerosis. N. Engl. J. Med., 326, 1464–8CrossRefGoogle ScholarPubMed
Rothstein, J. D., Kammen, M., Levey, A. I., Martin, L. J., and Kuncl, R. W. (1995). Selective loss of glial glutamate transporter GLT-1 in amyotrophic lateral sclerosis. Ann. Neurol., 38, 73–84CrossRefGoogle ScholarPubMed
Rowland, L. P. and Shneider, N. A. (2001). Amyotrophic lateral sclerosis. N. Engl. J. Med., 344, 1688–700CrossRefGoogle ScholarPubMed
Roy, N., Mahadevan, M. S., McLean, M.et al. (1995). The gene for neuronal apoptosis inhibitory protein is partially deleted in individuals with spinal muscular atrophy. Cell, 80, 167–78CrossRefGoogle ScholarPubMed
Sakahira, H., Enari, M., Ohsawa, Y., Uchiyama, Y., and Nagata, S. (1999). Apoptotic nuclear morphological change without DNA fragmentation. Curr. Biol., 9, 543–6CrossRefGoogle ScholarPubMed
Sano, T., Lin, H., Chen, X.et al. (1999). Differential expression of MMAC/PTEN in glioblastoma multiforme: relationship to localization and prognosis. Cancer Res., 59, 1820–4Google ScholarPubMed
Sathasivam, S., Ince, P. G., and Shaw, P. J. (2001). Apoptosis in amyotrophic lateral sclerosis: a review of the evidence. Neuropathol. Appl. Neurobiol., 27, 257–74CrossRefGoogle Scholar
Satoh, T., Nakatsuka, D., Watanabe, Y., Nagata, I., Kikuchi, H., and Namura, S. (2000). Neuroprotection by MAPK/ERK kinase inhibition with U0126 against oxidative stress in a mouse neuronal cell line and rat primary cultured cortical neurons. Neurosci. Lett., 288, 163–6CrossRefGoogle Scholar
Saudou, F., Finkbeiner, S., Devys, D., and Greenberg, M. E. (1998). Huntingtin acts in the nucleus to induce apoptosis but death does not correlate with the formation of intranuclear inclusions. Cell, 95, 55–66CrossRefGoogle Scholar
Saunders, J. W. (1966). Death in embryonic systems. Science, 154, 604–12CrossRefGoogle ScholarPubMed
Schilling, G., Becher, M. W., Sharp, A. H.et al. (1999). Intranuclear inclusions and neuritic aggregates in transgenic mice expressing a mutant N-terminal fragment of huntingtin. Hum. Mol. Genet., 8, 397–407CrossRefGoogle ScholarPubMed
Schreiber, S. S., Tocco, G., Najm, I., Thompson, R. F., and Baudry, M. (1993). Cycloheximide prevents kainate-induced neuronal death and c-fos expression in adult rat brain. J. Mol. Neurosci., 4, 149–59CrossRefGoogle ScholarPubMed
Schwartz, L. M. and Milligan, C. E. (1996). Cold thoughts of death: the role of ICE proteases in neuronal cell death. Trends Neurosci., 19, 555–62CrossRefGoogle ScholarPubMed
Schwartz, L. M., Smith, S. W., Jones, M. E. E. M, and Osborne, B. A. (1993). Do all programmed cell deaths occur via apoptosis?Proc. Natl. Acad. Sci. USA, 90, 980–4CrossRefGoogle ScholarPubMed
Schweichel, J. U. and Merker, H. J. (1973). The morphology of various types of cell death in prenatal tissues. Teratology, 7, 253–66CrossRefGoogle ScholarPubMed
Scorrano, L., Oakes, S. A., Opferman, T. J.et al. (2003). Bax and Bak regulation of endoplasmic reticulum Ca2+: a control point for apoptosis. Science, 300, 135–9CrossRefGoogle ScholarPubMed
Selkoe, D. J. (2002). Alzheimer's disease is a synaptic failure. Science, 298, 789–91CrossRefGoogle ScholarPubMed
Selznick, L. A., Holtzman, D. M., Han, B. H.et al. (1999). In situ immunodetection of neuronal caspase-3 activation in Alzheimer's disease. J. Neuropath. Exp. Neurol., 58, 1020–6CrossRefGoogle Scholar
Shaikh, A. Y. and Martin, L. J. (2002). DNA base-excision repair enzyme apurinic/apyrimidinic endonuclease/redox factor-1 is increased and competent in brain and spinal cord of individuals with amyotrophic lateral sclerosis. NeuroMolecular Med., 2, 47–60CrossRefGoogle ScholarPubMed
Sherrington, R., Rogaev, E. I., Liang, Y.et al. (1995). Cloning of a gene bearing missense mutations in early-onset familial Alzheimer's disease. Nature, 375, 754–60CrossRefGoogle ScholarPubMed
Shieh, S.-Y., Ikeda, M., Taya, Y., and Prives, C. (1997). DNA damage-induced phosphorylation of p53 alleviates inhibition by MDM2. Cell, 91, 325–34CrossRefGoogle ScholarPubMed
Shimizu, S., Ide, T., Yanagida, T., and Tsujimoto, Y. (2000). Electrophysiological study of a novel large pore formed by Bax and the voltage-dependent anion channel that is permeable to cytochrome c. J. Biol. Chem., 275, 12321–5CrossRefGoogle ScholarPubMed
Shimohama, S., Tanino, H., and Fujimoto, S. (2001). Differential subcellular localization of caspase family protein in the adult rat brain. Neurosci. Lett., 315, 125–8CrossRefGoogle ScholarPubMed
Simonian, N. A., Getz, R. L., Leveque, J. C., Konradi, C., and Coyle, J. T. (1996). Kainate induces apoptosis in neurons. Neuroscience, 74, 675–83CrossRefGoogle ScholarPubMed
Simons, M., Beinroth, S., Gleichmann, M.et al. (1999). Adenovirus-mediated gene transfer of inhibitors of apoptosis proteins delays apoptosis in cerebellar granule neurons. J. Neurochem., 72, 292–301CrossRefGoogle ScholarPubMed
Smale, G., Nichols, N. R., Brady, D. R., Finch, C. E., and Horten, W. E. Jr. (1995). Evidence for apoptotic cell death in Alzheimer's disease. Exp. Neurol., 133, 225–30CrossRefGoogle ScholarPubMed
Song, Q., Kuang, Y., Dixit, V. M., and Vincenz, C. (1999). Boo, a negative regulator of cell death, interacts with Apaf-1. EMBO J., 18, 167–78CrossRefGoogle ScholarPubMed
Srivastava, R. K., Sollott, S. J., Khan, L., Hansford, R., Lakatta, E. G., and Longo, D. L. (1999). Bcl-2 and Bcl-X(L) block thapsigargin-induced nitric oxide generation, c-Jun NH(2)-terminal kinase activity, and apoptosis. Mol. Cell Biol., 19, 5659–74CrossRefGoogle ScholarPubMed
Stadelmann, C., Deckwerth, T. L., Srinivasan, A., Bancher, C., Brock, W., and Lassmann, H. (1999). Activation of caspase-3 in single neurons and autophagic granules of granulovacuolar degeneration in Alzheimer's disease. Evidence for apoptotic cell death. Am. J. Pathol., 155, 1459–66CrossRefGoogle ScholarPubMed
Stanciu, M., Wang, Y., Kentor, R.et al. (2000). Persistent activation of ERK contributes to glutamate-induced oxidative toxicity in a neuronal cell line and primary cortical neuron cultures. J. Biol. Chem., 275, 12200–6CrossRefGoogle Scholar
Steffan, J. S., Kazantsev, A., Spasic-Boskovic, O.et al. (2000). The Huntington's disease protein interacts with p53 and CREB-binding protein and represses transcription. Proc. Natl. Acad. Sci. USA, 97, 6763–8CrossRefGoogle ScholarPubMed
Stennicke, H. R., Deveraux, Q. L., Humke, E. W., Reed, J. C., Dixit, V. M., and Salvesen, G. S. (1999). Caspase-9 can be activated without proteolytic processing. J. Biol. Chem., 274, 8359–62CrossRefGoogle ScholarPubMed
Su, J. H., Anderson, A. J., Cribbs, D. H.et al. (2003). Fas and Fas ligand are associated with neuritic degeneration in the AD brain and participate in β-amyloid-induced neuronal death. Neurobiol. Dis., 12, 182–93CrossRefGoogle ScholarPubMed
Su, J. H., Zhao, M., Anderson, A. J., Srinivasan, A., and Cotman, C. W. (2001). Activated caspase-3 expression in Alzheimer's and aged control brain: correlation with Alzheimer pathology. Brain Res., 898, 350–7CrossRefGoogle ScholarPubMed
Susin, S. A., Lorenzo, H. K., Zamzami, N.et al. (1999). Molecular characterization of mitochondrial apoptosis-inducing factor. Nature, 397, 441–6CrossRefGoogle ScholarPubMed
Sze, C.-I., Bi, H., Kleinschmidt-DeMasters, B. K., Filley, C. M., and Martin, L. J. (2001). N-Methyl-D-aspartate receptor subunit proteins and their phosphorylation status are altered selectively in Alzheimer's disease. J. Neurol. Sci., 182, 151–9CrossRefGoogle ScholarPubMed
Sze, C.-I., Troncoso, J. C., Kawas, C., Mouton, P., Price, D. L., and Martin, L. J. (1997). Loss of the presynaptic vesicle protein synaptophysin in hippocampus correlates with cognitive decline in Alzheimer's disease. J. Neuropathol. Exp. Neurol., 56, 933–94CrossRefGoogle Scholar
Takashima, A., Noguchi, K., Sato, K., Hoshino, T., and Imahori, K . (1993). Tau protein kinase I is essential for amyloid beta-protein-induced neurotoxicity. Proc. Natl. Acad. Sci. USA, 90, 7789–93CrossRefGoogle ScholarPubMed
Tandan, R., Robison, S. H., Munzer, J. S. and Bradley, W. G. (1987). Deficient DNA repair in amyotrophic lateral sclerosis cells. J. Neurol. Sci., 79, 189–203CrossRefGoogle ScholarPubMed
Tata, J. R. (1966). Requirement for RNA and protein synthesis for induced regression of tadpole tail in organ culture. Dev. Biol., 13, 77–94CrossRefGoogle ScholarPubMed
Tatton, N. A., MaClean-Fraser, A., Tatton, W. G., Perl, D. P., and Olnanow, C. W. (1998). A fluorescent double-labeling method to detect and confirm apoptotic nuclei in Parkinson's disease. Ann. Neurol., 44, S142–8CrossRefGoogle ScholarPubMed
Tatton, W. G., Chalmers-Redman, R., Brown, D., and Tatton, N. (2003). Apoptosis in Parkinson's disease: signals for neuronal degeneration. Ann. Neurol., 53, S61–72CrossRefGoogle Scholar
Tenneti, L. and Lipton, S. A. (2000). Involvement of activated caspase-3-like proteases in N-methyl-D-aspartate-induced apoptosis in cerebrocortical neurons. Ann. Neurol., 74, 134–42Google ScholarPubMed
Terry, R. D., Masliah, E., Salmon, D. P.et al. (1991). Physical basis of cognitive alterations in Alzheimer's disease: synapse loss is the major correlate of cognitive impairment. Ann. Neurol., 30, 572–80CrossRefGoogle ScholarPubMed
Tomkins, J., Dempster, S., Banner, S. J., Cookson, M. R., and Shaw, P. J. (2000). Screening of AP endonuclease as a candidate gene for amyotrophic lateral sclerosis. NeuroReport, 11, 1695–7CrossRefGoogle ScholarPubMed
Tompkins, M. M., Basgall, E. J., Zamrini, E., and Hill, W. D. (1997). Apoptotic-like changes in Lewy-body-associated disorders and normal aging in substantia nigra neurons. Am. J. Pathol., 150, 119–31Google Scholar
Torvik, A. (1976). Central chromatolysis and the axon reaction. A reappraisal. Neuropathol. Appl. Neurobiol., 2, 423–32CrossRefGoogle Scholar
Trimmer, P. A., Smith, T. S., Jung, A. B., and Bennett, J. P. Jr. (1996). Dopamine neurons from transgenic mice with a knockout of the p53 gene resist MPTP neurotoxicity. Neurodegeneration, 5, 233–9CrossRefGoogle ScholarPubMed
Troy, C. M., Friedman, J. E., and Friedman, W. J. (2002). Mechanisms of p75-mediated death of hippocampal neurons. Role of caspases. J. Biol. Chem., 277, 34295–302CrossRefGoogle ScholarPubMed
Trump, B. F. and Berezesky, I. K. (1996). The role of altered [Ca2+]i regulation in apoptosis, oncosis, and necrosis. Biochim. Biophys. Acta., 1313, 173–8CrossRefGoogle ScholarPubMed
Trump, B. J., Goldblatt, P. J., and Stowell, R. E. (1964). Studies on necrosis of mouse liver in vitro. Ultrastructural alterations in the mitochondria of hepatic parenchymal cells. Lab. Invest., 14, 343–71Google Scholar
Turmaine, M., Raza, A., Mahal, A., Mangiarini, L., Bates, G. P., and Davies, S. W. (2000). Nonapoptotic neurodegeneration in a transgenic mouse model of Huntington's disease. Proc. Natl. Acad. Sci. USA, 97, 8093–7CrossRefGoogle Scholar
Uberti, D., Carsana, T., Bernardi, E.et al. (2002). Selective impairment of p53-mediated cell death in fibroblasts from sporadic Alzheimer's disease patients. J. Cell Sci., 115, 3131–8Google ScholarPubMed
Uetsuki, T., Takemoto, K., Nishimura, I.et al. (1999). Activation of neuronal caspase-3 by intracellular accumulation of wild-type Alzheimer amyloid precursor protein. J. Neurosci., 19, 6955–64CrossRefGoogle ScholarPubMed
Uppuluri, S., Knipling, L., Sackett, D. L., and Wolff, J. (1993). Localization of the colchicine-binding site of tubulin. Proc. Natl. Acad. Sci. USA, 90, 11598–602CrossRefGoogle ScholarPubMed
Bekkum, D. W. (1957). The effect of x-rays on phosphorylations in vivo. Biochim. Biophys. Acta., 25, 487–92CrossRefGoogle Scholar
Vander Heiden, M. G., Chandel, N. S., Williamson, E. K., Schumacker, P. T., and Thompson, C. B. (1997). Bcl-xL regulates the membrane potential and volume homeostasis of mitochondria. Cell, 91, 627–37CrossRefGoogle ScholarPubMed
Gurp, M., Festjens, N., Loo, G., Saelens, X., and Vandenabeele, P. (2003). Mitochondrial intermembrane proteins in cell death. Biochem. Biophys. Res. Comm., 304, 487–97CrossRefGoogle ScholarPubMed
Lookeren Campagne, M., Lucassen, P. J., Vermeulen, J. P., and Balázs, R. (1995). NMDA and kainate induced internucleosomal DNA cleavage associated with both apoptotic and necrotic cell death in the neonatal rat brain. Eur. J. Neurosci., 7, 1627–40CrossRefGoogle ScholarPubMed
Verhagen, A. M., Ekert, P. G., Pakusch, M.et al. (2000). Identification of DIABLO, a mammalian protein that promotes apoptosis by binding to and antagonizing IAP proteins. Cell, 102, 43–53CrossRefGoogle ScholarPubMed
Verhagen, A. M., Silke, J., Ekert, P. G.et al. (2002). HtrA2 promotes cell death through its serine protease activity and its ability to antagonize inhibitor of apoptosis proteins. J. Biol. Chem., 277, 445–54CrossRefGoogle ScholarPubMed
Vila, M., Jackson-Lewis, V., Vukosavic, S.et al. (2001). Bax ablation prevents dopaminergic neurodegeneration in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine mouse model of Parkinson's disease. Proc. Natl. Acad. Sci., 98, 2837–42CrossRefGoogle ScholarPubMed
Villunger, A., Michalak, E. M., Coultas, L.et al. (2003). p53- and drug-induced apoptotic responses mediated by BH3-only proteins Puma and Noxa. Science, 302, 1036–8CrossRefGoogle ScholarPubMed
Virchow, R. (1858). Cellular Pathology as Based Upon Physiological and Pathological Histology (Translation). London: Churchill
Vonsattel, J. P. and DiFiglia, M. (1998). Huntington's disease. J. Neuropathol. Exp. Neurol., 57, 369–84CrossRefGoogle Scholar
Vonsattel, J. P., Myers, R. H., Stevens, T. J., Ferrante, R. J., Bird, E. D., and Richardson, E. P. Jr. (1985). Neuropathological classification of Huntington's disease. J. Neuropathol. Exp. Neurol., 44, 559–77CrossRefGoogle ScholarPubMed
Vukosavic, S., Dubois-Dauphin, N., Romero, N., and Przedborski, S. (1999). Bax and Bcl-2 interaction in a transgenic mouse model of amyotrophic lateral sclerosis. J. Neurochem., 73, 2460–8CrossRefGoogle Scholar
Wang, H.-G., Pathan, N., Ethell, I. M.et al. (1999). Ca2+-induced apoptosis through calcineurin dephosphorylation of Bad. Science, 284, 339–43CrossRefGoogle ScholarPubMed
Wang, H.-G., Rapp, U. R., and Reed, J. C. (1996). Bcl-2 targets the protein kinase raf-1 to mitochondria. Cell, 87, 629–38CrossRefGoogle Scholar
Wei, M. C., Zong, W-X., Cheng, E. H.-Y.et al. (2001). Proapoptotic Bax and Bak: a requisite gateway to mitochondrial dysfunction and death. Science, 292, 727–30CrossRefGoogle ScholarPubMed
Weidemann, A., Paliga, K., Drrwang, U.et al. (1999). Proteolytic processing of the Alzheimer's disease amyloid precursor protein within its cytoplasmic domain by caspase-like proteases. J. Biol. Chem., 274, 5823–9CrossRefGoogle ScholarPubMed
West, M. J., Kawas, C. H., Martin, L. J., and Troncoso, J. C. (2000). The CA1 region of the human hippocampus is a hot spot in Alzheimer's disease. Ann. NY Acad. Sci., 908, 255–9CrossRefGoogle ScholarPubMed
White, E. and Prives, C. (1999). DNA damage enables p73. Nature, 399, 734–7CrossRefGoogle ScholarPubMed
Whitehouse, P. J., Price, D. L., Struble, R. G., Clark, A. W., Coyle, J. T., and DeLong, M. R. (1982). Alzheimer's disease and senile dementia: loss of neurons in the basal forebrain. Science, 215, 1237–9CrossRefGoogle ScholarPubMed
Widmann, C., Gerwins, P., Johnson, N., Jarpe, M. B., and Johnson, G. L. (1998). MEK kinase 1, a substrate for DEVD-directed caspases, is involved in genotoxin-induced apoptosis. Mol. Cell Biol., 18, 2416–29CrossRefGoogle ScholarPubMed
Wolf, B. B. and Green, D. R. (1999). Suicidal tendencies: apoptotic cell death by caspase family proteinases. J. Biol. Chem., 274, 20049–52CrossRefGoogle ScholarPubMed
Wolter, K. G., Hsu, Y.-T., Smith, C. L., Nechushtan, A., Xi, X.-G., and Youle, R. L. (1997). Movement of Bax from the cytosol to mitochondria during apoptosis. J. Cell Biol., 139, 1281–92CrossRefGoogle ScholarPubMed
Wong, P. C., Cai, H., Borchelt, D. R., and Price, D. L. (2002). Genetically engineered mouse models of neurodegenerative diseases. Nat. Neurosci., 5, 633–9CrossRefGoogle ScholarPubMed
Wong, P. C., Pardo, C. A., Borchelt, D. R.et al. (1995). An adverse property of a familial ALS-linked SOD1 mutation causes motor neuron disease characterized by vacuolar degeneration of mitochondria. Neuron, 14, 1105–16CrossRefGoogle ScholarPubMed
Wüllner, U., Kornhuber, J., Weller, M.et al. (1999). Cell death and apoptosis regulating proteins in Parkinson's disease – a cautionary note. Acta. Neuropathol., 97, 408–12Google ScholarPubMed
Wyllie, A. H. (1980). Glucocorticoid-induced thymocyte apoptosis is associated with endogenous endonuclease activation. Nature, 284, 555–6CrossRefGoogle ScholarPubMed
Wyllie, A. H., Kerr, J. F. R., and Currie, A. R. (1980). Cell death: the significance of apoptosis. Int. Rev. Cytol., 68, 251–306CrossRefGoogle Scholar
Xanthoudakis, S. and Curran, T. (1992). Identification and characterization of Ref-1, a nuclear protein that facilitates AP-1 DNA-binding activity. EMBO J., 11, 653–65Google ScholarPubMed
Xiang, H., Kinoshita, Y., Knudson, C. M., Korsmeyer, S. J., Schwartzkroin, P. A., and Morrison, R. S. (1998). Bax involvement in p53-mediated neuronal cell death. J. Neurosci., 18, 1363–73CrossRefGoogle ScholarPubMed
Xu, D. G., Korneluk, R. G., Tamai, K.et al. (1997). Distribution of neuronal apoptosis inhibitory protein-like immunoreactivity in the rat central nervous system. J. Comp. Neurol., 382, 247–593.0.CO;2-3>CrossRefGoogle ScholarPubMed
Xu, J., Kao, S. Y., Lee, F. J., Song, W., Jin, L. W., and Yankner, B. A. (2002). Dopamine-dependent neurotoxicity of alpha-synuclein: a mechanism for selective neurodegeneration in Parkinson disease. Nat. Med., 8, 600–6CrossRefGoogle ScholarPubMed
Xue, L. Z., Fletcher, G. C., and Tolkovsky, A. M. (1999). Autophagy is activated by apoptotic signalling in sympathetic neurons: an alternative mechanism of death execution. Mol. Cell Neurosci., 14, 180–98CrossRefGoogle ScholarPubMed
Yaar, M., Zhai, S., Fine, R. E.et al. (2002). Amyloid β binds trimers as well as monomers of the 75-kDa neurotrophin receptor and activates receptor signaling. J. Biol. Chem., 277, 7720–5CrossRefGoogle ScholarPubMed
Yan, X. X., Najbauer, J., Woo, C. C., Dashtipour, K., Ribak, C. E., and Leon, M. (2001). Expression of active caspase-3 in mitotic and postmitotic cells or rat forebrain. J. Comp. Neurol., 433, 4–22CrossRefGoogle ScholarPubMed
Yang, A., Walker, N., Bronson, R.et al. (2000). p73-deficient mice have neurological, pheromonal and inflammatory defects but lack spontaneous tumors. Nature, 404, 99–103CrossRefGoogle Scholar
Yang, E., Zha, J., Jockel, J., Boise, L. H., Thompson, C. B., and Korsmeyer, S. J. (1995). Bad, a heterodimeric partner for Bcl-xL and Bcl-2, displaces Bax and promotes cell death. Cell, 80, 285–91CrossRefGoogle ScholarPubMed
Yang, J., Liu, X., Bhalla, K.et al. (1997). Prevention of apoptosis by bcl-2: release of cytochrome c from mitochondria blocked. Science, 275, 1129–32CrossRefGoogle ScholarPubMed
Yang, L., Matthews, R. T., Schulz, J. B.et al. (1998). 1-methyl-4-phenyl-1,2,3,6-tetrahydropyride neurotoxicity is attenuated in mice overexpressing Bcl-2. J. Neurosci., 18, 8145–52CrossRefGoogle ScholarPubMed
Yankner, B. A., Dawes, L. R., Fisher, S.et al. (1989). Neurotoxicity of a fragment of the amyloid precursor associated with Alzheimer's disease. Science, 245, 417–20CrossRefGoogle ScholarPubMed
Yoshio, H., Kong, Y.-Y., Yoshida, R.et al. (1998). Apaf1 is required for mitochondrial pathways of apoptosis and brain development. Cell, 94, 739–50CrossRefGoogle Scholar
Younkin, S. G. (1995). Evidence that A beta 42 is the real culprit in Alzheimer's disease. Ann. Neurol., 37, 287–8CrossRefGoogle ScholarPubMed
Yu, Z. X., Li, S. H., Evans, J., Pillarsetti, A., Li, H., and Li, X. J. (2003). Mutant huntingtin causes context – dependent neurodegeneration in mice with Huntington's disease. J. Neurosci., 23, 2193–202CrossRefGoogle ScholarPubMed
Yuan, J., Lipinski, M., and Degterev, A. (2003). Diversity in the mechanisms of neuronal cell death. Neuron, 40, 401–13CrossRefGoogle ScholarPubMed
Yue, Z., Horton, A., Bravin, M., DeJager, P. L., Selimi, F., and Heintz, N. (2002). A novel protein complex linking the δ2 glutamate receptor and autophagy: implications for neurodegeneration in Lurcher mice. Neuron, 35, 921–33CrossRefGoogle Scholar
Zha, J., Harada, H., Yang, E., Jockel, J., and Korsmeyer, S. J. (1996). Serine phosphorylation of death agonist Bad in response to survival factor results in binding to 14-3-3 not Bcl-xL. Cell, 87, 619–28CrossRefGoogle ScholarPubMed
Zhang, Y., McLaughlin, R., Goodyer, C., and LeBlanc, A. (2002). Selective cytotoxicity of intracellular amyloid β peptide1-42 through p53 and Bax in cultured primary human neurons. J. Cell Biol., 156, 519–29CrossRefGoogle ScholarPubMed
Zong, W.-X., Li, C., Hatzivassiliou, G.et al. (2003). Bax and Bak can localize to the endoplasmic reticulum to initiate apoptosis. J. Cell Biol., 162, 59–69CrossRefGoogle ScholarPubMed
Zou, H., Li, Y., Liu, X., and Wang, X. (1999). An Apaf-1-cytochrome c multimeric complex is a functional apoptosome that activates procaspase-9. J. Biol. Chem., 274, 11549–56CrossRefGoogle ScholarPubMed
Zuch, C. L., Nordstroem, V. K., Briedrick, L. A., Hoering, G. R., Granholm, A.-C., and Bickford, P. C. (2000). Time course of degenerative alterations in nigral dopaminergic neurons following a 6-hydroxydopamine lesion. J. Comp. Neurol., 427, 440–543.0.CO;2-7>CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Neuronal cell death in human neurodegenerative diseases and their animal/cell models
    • By Lee J. Martin, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Zhiping Liu, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Juan Troncoso, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Donald L. Price, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA
  • Edited by Martin Holcik, University of Ottawa, Eric C. LaCasse, University of Ottawa, Alex E. MacKenzie, University of Ottawa, Robert G. Korneluk, University of Ottawa
  • Book: Apoptosis in Health and Disease
  • Online publication: 03 March 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511663543.005
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Neuronal cell death in human neurodegenerative diseases and their animal/cell models
    • By Lee J. Martin, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Zhiping Liu, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Juan Troncoso, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Donald L. Price, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA
  • Edited by Martin Holcik, University of Ottawa, Eric C. LaCasse, University of Ottawa, Alex E. MacKenzie, University of Ottawa, Robert G. Korneluk, University of Ottawa
  • Book: Apoptosis in Health and Disease
  • Online publication: 03 March 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511663543.005
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Neuronal cell death in human neurodegenerative diseases and their animal/cell models
    • By Lee J. Martin, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Zhiping Liu, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Juan Troncoso, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA, Donald L. Price, Departments of Pathology, Division of Neuropathology, Neuroscience, and Neurology, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA
  • Edited by Martin Holcik, University of Ottawa, Eric C. LaCasse, University of Ottawa, Alex E. MacKenzie, University of Ottawa, Robert G. Korneluk, University of Ottawa
  • Book: Apoptosis in Health and Disease
  • Online publication: 03 March 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511663543.005
Available formats
×