Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-29T03:06:07.226Z Has data issue: false hasContentIssue false

3 - Mitochondria, metabolic inhibitors and neurodegeneration

from Part I - Basic aspects of neurodegeneration

Published online by Cambridge University Press:  04 August 2010

M. Flint Beal
Affiliation:
Cornell University, New York
Anthony E. Lang
Affiliation:
University of Toronto
Albert C. Ludolph
Affiliation:
Universität Ulm, Germany
James G. Greene
Affiliation:
Center for Neurodegenerative Disease and Department of Neurology, Emory University, Atlanta, GA, USA
J. Timothy Greenamyre
Affiliation:
Pittsburgh Institute for Neurodegenerative Diseases, University of Pittsburgh, PA, USA
Get access

Summary

The role of the mitochondrion in neurodegeneration is a paradox. On the one hand, vital mitochondrial tasks, such as energy production and calcium buffering, provide an important foundation for all neuronal functions. Yet, on the other, mitochondrial free radical production and involvement in cell-death cascades may lead to a neuron's untimely demise.

It is now clear that mitochondria are not merely neuronal “power plants”, but are highly complex, integrated organelles whose function transcends that of simple energy production. In addition to providing the majority of neuronal energy via oxidative phosphorylation, mitochondria play a central role in intracellular ion homeostasis, free radical management, and gene and protein expression.

This chapter will focus on the biology of mitochondrial electron transport, oxidative phosphorylation and other mitochondrial functions, and will discuss the effects of mitochondrial toxins on mitochondrial function and neuronal viability. It will explore briefly one of the main consequences of oxidative metabolism, mitochondrial free radical production and how this and other mitochondrial factors potentially contribute to neuronal death.

Mitochondrial energy production and sites of action for metabolic inhibitors

Mitochondria efficiently convert the potential energy of glucose into a usable cellular energy currency, primarily ATP. Glucose is the primary basis for neuronal energy metabolism; ketone bodies can provide a limited energy source, but only in situations of chronic metabolic imbalance.

Glucose crosses the blood–brain barrier in an insulin-independent manner and is taken up by membrane transporters. It is phosphorylated almost immediately by hexokinase and enters glycolysis.

Type
Chapter
Information
Neurodegenerative Diseases
Neurobiology, Pathogenesis and Therapeutics
, pp. 33 - 43
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

Alston, T. A., Mela, L. & Bright, H. J. (1977). 3-nitropropionate, the toxic substance of Indigofera, is a suicide inactivator of succinate dehydrogenase. Proc. Natl Acad. Sci., USA, 74, 3767–71CrossRefGoogle ScholarPubMed
Andreassen, O. A., Ferrante, R. J. & Dedeoglu, A. (2001). Mice with a partial deficiency of manganese superoxide dismutase show increased vulnerability to the mitochondrial toxins malonate, 3-nitropropionic acid, and MPTP. Exp. Neurol., 167, 189–95CrossRefGoogle ScholarPubMed
Andreeva, N., Khodorov, B., Stelmashook, E.et al. (1991). Inhibition of Na/Ca exchange enhances delayed neuronal deathelicited by glutamate in cerebellar granule cell cultures. Brain Res., 548, 322–5CrossRefGoogle Scholar
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., Montessuit, S., Sanchez, B. & Martinou, J. C. (2001). Bax is present as a high molecular weight oligomer/complex in the mitochondrial membrane of apoptotic cells. J. Biol. Chem., 276, 11615–23CrossRefGoogle ScholarPubMed
Astrup, J., Sorensen, P. M. & Sorensen, H. R. (1981). Oxygen and glucose consumption related to Na–K transport in canine brain. Stroke, 12, 726–30CrossRefGoogle ScholarPubMed
Beal, M. F. (1998). Mitochondrial dysfunction in neurodegenerative diseases. Biochim. Biophys. Act., 1366, 211–23CrossRef
Beal, M. F., Ferrante, R. J., Henshaw, R.et al. (1993). Neurochemical and histologic characterization of striatal excitotoxic lesions produced by the mitochondrial toxin 3-nitropropionic acid. J. Neurosci., 13, 4181–92CrossRefGoogle ScholarPubMed
Beal, M. F., Ferrante, R. J., Henshaw, R.et al. (1995). 3-Nitropropionic acid neurotoxicity is attenuated in copper/zinc superoxide dismutase transgenic mice. J. Neurochem., 65, 919–22CrossRefGoogle ScholarPubMed
Bennett, M. C., Mlady, G. W., Fleshner, M. & Rose, G. M. (1996). Synergy between chronic corticosterone and sodium azide treatments in producing a spatial learning deficit and inhibiting cytochrome oxidase activity. Proc. Natl Acad, Sci., USA, 93, 1330–4CrossRefGoogle ScholarPubMed
Bernardi, P., Broekemeier, K. M. & Pfeiffer, D. R. (1994). Recent progress on regulation of the mitochondrial permeability transition pore: a cyclosperin-sensitive pore in the inner mitochondrial membrane. J. Bioenerg. Biomembr., 26, 509–17CrossRefGoogle ScholarPubMed
Betarbet, R., Sherer, T. B., Mackenzie, G., Garcia-Osuna, M., Panov, A. V. & Greenamyre, J. T. (2000). Chronic systemic pesticide exposure reproduces features of Parkinson's disease. Nat. Neurosc., 3, 1301–6CrossRefGoogle ScholarPubMed
Blaustein, M. P. (1988). Calcium transport and buffering in neurons. Trends Neurosci., 10, 438–43CrossRefGoogle Scholar
Bogdanov, M. B., Ferrante, R. J., Mueller, G., Ramos, L. E., Martinou, J. C. & Beal, M. F. (1999). Oxidative stress is attenuated in mice overexpressing BCL-2. Neurosci. Lett., 262, 33–6CrossRefGoogle ScholarPubMed
Boveris, A. & Turrens, J. F. (1980). Production of superoxide anion by the NADH-dehydrogenase of mammalian mitochondria. In Aspects in Superoxide and Superoxide Dismutase. Developments in Biochemistry, ed. J. V. Bannister and H. A. O. Hill, Vol. 11A, pp. 84–91. New York: Elsevier-North Holland
Boveris, A., Oshino, N. & Chance, B. (1972). The cellular production of hydrogen peroxide. Biochem. J., 128, 617–30CrossRefGoogle ScholarPubMed
Brouillet, E., Hyman, B. T., Jenkins, B. G.et al. (1994). Systemic or local administration of azide produces striatal lesions by an energy impairment-induced excitotoxic mechanism. Exp. Neurol., 129, 175–82CrossRefGoogle ScholarPubMed
Brouillet, E., Hantraye, P., Ferrante, R. J.et al. (1995). Chronic mitochondrial energy impairment produces selective striatal degeneration and abnormal choreiform movements in primates. Proc. Natl Acad. Sci., USA, 92, 7105–9CrossRefGoogle ScholarPubMed
Cao, X. & Phillis, J. W. (1994). Alpha-phenyl-test-butyl nitrone reduces cortical infarct and edema in rats subjected to focal ischemia.Brain Res., 644, 267–72CrossRefGoogle ScholarPubMed
Carter, A. J., Muller, R. E., Pschorn, U. & Stransky, W. (1995). Preincubation with creatine enhances levels of creatine phosphate and prevents anoxic damage in rat hippocampal slices. J. Neurochem., 64, 2691–9CrossRefGoogle ScholarPubMed
Cousin, M. A., Nicholls, D. G. & Pocock, J. M. (1995). Modulation of ion gradients and glutamate release in cultured cerebellar granule cells by ouabain. J. Neurochem., 64, 2097–104CrossRefGoogle ScholarPubMed
Crompton, M. (1999). The mitochondrial permeability transition pore and its role in cell death. Biochem. J., 341, 233–49CrossRefGoogle ScholarPubMed
Crompton, M., Virji, S. & Ward, J. M. (1998). Cyclophilin-D binds strongly to complexes of the voltage-dependent anion channel and the adenine nucleotide translocase to form the permeability transition pore. Eur. J. Biochem., 258, 729–35CrossRefGoogle Scholar
Davey, G. P., Peuchen, S. & Clark, J. B. (1998). Energy thresholds in brain mitochondria. Potential involvement in neurodegeneration. J. Biol. Chem., 273, 12753–7CrossRefGoogle ScholarPubMed
Ding, H. F. & Fisher, D. E. (1998). Mechanisms of p53-mediated apoptosis. Crit. Rev. Oncol., 9, 83–98CrossRefGoogle ScholarPubMed
Du, C., Fang, M., Li, Y. & Wang, X. (2000). Smac, a mitochondrial protein that promotes cytochrome c-dependent caspase activation by eliminating IAP inhibition. Cell, 102, 33–42CrossRefGoogle ScholarPubMed
Ekert, P. G., Silke, J., Hawkins, C. J., Verhagen, A. M. & Vaux, D. L. (2001). DIABLO promotes apoptosis by removing MIHA/XIAP from processed caspase 9. J. Cell Biol., 152, 483–90CrossRefGoogle ScholarPubMed
Erecinska, M. & Dagani, F. (1990) Relationships between the neuronal sodium/potassium pump and energy metabolism. J. Gen. Physiol., 95, 591–616CrossRefGoogle ScholarPubMed
Fiskum, G., Murphy, A. N. & Beal, M. F. (1999). Mitochondria in neurodegeneration: acute ischemia and chronic neurodegenerative diseases. J. Cereb. Blood Flow Metab., 19, 351–69CrossRefGoogle ScholarPubMed
Flint, D. H., Tuminello, J. F. & Emptage, M. H. (1993). The inactivation of Fe-S cluster containing hydro-lases by superoxide. J. Biol. Chem., 268, 22369–76Google Scholar
Fridovitch, I. (1995). Superoxide radical and superoxide dismutases. Ann. Rev. Biochem., 64, 97–112CrossRefGoogle Scholar
Greenamyre, J. T., Garcia-Osuna, M. & Greene, J. G. (1994). The endogenous cofactors, thioctic acid and dihydrolipoic acid, are neuroprotective against NMDA and malonic acid lesions of striatum. Neurosci. Lett., 171, 17–20CrossRefGoogle ScholarPubMed
Greene, J. G. & Greenamyre, J. T. (1995). Characterization of the excitotoxic potential of the reversible succinate dehydrogenase inhibitor malonate. J. Neurochem., 64, 430–6CrossRefGoogle ScholarPubMed
Greene, J. G. & Greenamyre, J. T. (1996). Manipulation of membrane potential modulates malonate-induced striatal excitotoxicity. J. Neurochem., 66, 637–43CrossRefGoogle ScholarPubMed
Greene, J. G., Sheu, S. S., Gross, R. A. & Greenamyre, J. T. (1998). 3-nitropropionic acid exacerbates N-methyl-D-aspartate toxicity in striatal culture by multiple mechanisms. Neuroscience, 84, 503–10CrossRefGoogle ScholarPubMed
Gunasekar, P. G., Kanthasamy, A. G., Borowitz, J. L. & Isom, G. (1995). NMDA receptor activation produces concurrent generation of nitric oxide and reactive oxygen species: implication for cell death. J. Neurochem., 65, 2016–21CrossRefGoogle ScholarPubMed
Gunter, T. E. & Gunter, K. K. (2001). Uptake of calcium by mitochondria: transport and possible function. IUBMB Life, 52, 197–204CrossRefGoogle ScholarPubMed
Gunter, T. E., Gunter, K. K., Sheu, S.- S. & Gavin, C. E. (1994). Mitochondrial calcium transport. Am. J. Physiol., 267, C313–39CrossRefGoogle ScholarPubMed
Hengartner, M. O. (2000). The biochemistry of apoptosis. Nature, 407, 770–6CrossRefGoogle Scholar
Hensley, K., Pye, Q. N., Maidt, M.et al. (1998). Interaction of alpha-phenyl-N-tert-butyl nitrone and alternative electron acceptors with complex I indicates a substrate reduction site upstream from the rotenone binding site. J. Neurochem., 71, 2549–57CrossRefGoogle ScholarPubMed
Hochster R. M. & Quastel J. H., eds. (1963). Metabolic Inhibitors: A Comprehensive Treatise. New York: Academic Press
Ichas, F., Jouaville, L. S. & Mazat, J. P. (1997). Mitochondria are excitable organelles capable of generating and conveying electrical and calcium signals. Cell, 89, 1145–53CrossRefGoogle ScholarPubMed
Jenkins, B. G., Rosas, H. D., Chen, Y. C.et al. (1998). 1H NMR spectroscopy studies of Huntington's disease: correlations with CAG repeat numbers. Neurology, 50, 1357–65CrossRefGoogle ScholarPubMed
Kauppinen, R. A., Sihra, T. S. & Nicholls, D. G. (1987). Aminooxyacetic acid inhibits the malate–aspartate shuttle in isolated nerve terminals and prevents the mitochondria from using glycolytic substrates. Biochim. Biophys. Act., 930, 173–8CrossRefGoogle ScholarPubMed
Kerr, J. F., Wyllie, A. H. & Currie, A. R. (1972). Apoptosis: a basic biological phenomenon with wide ranging implications in tissue kinetics. Br. J. Cance., 26, 239–47CrossRefGoogle ScholarPubMed
Kiedrowski, L. & Costa, E. (1995). Glutamate-induced destabilization of intracellular calcium concentration homeostasis in cultured cerebellar granul cells: role of mitochondria in calcium buffering. Mol. Pharmacol., 53, 974–80Google Scholar
Kiedrowski, L., Brooker, G., Costa, E. & Wroblewski, J. T. (1994). Glutamate impairs neuronal calcium extrusion while reducing sodium gradient. Neuron, 12, 295–300CrossRefGoogle ScholarPubMed
Kirkland, R. A., Windelborn, J. A., Kasprzak, J. M. & Franklin, J. L. (2002). A Bax-induced pro-oxidant state is critical for cytochrome c release during programmed neuronal death. J. Neurosci., 22, 6480–90CrossRefGoogle ScholarPubMed
Koch, R. A. & Barrish, M. E. (1994). Perturbation of intracellular calcium and hydrogen ion regulation in cultured mouse hippocamp-al neurons by reduction of the sodium ion concentration gradient. J. Neurosci., 14, 2585–93CrossRefGoogle ScholarPubMed
Kroemer, G. & Reed, J. C. (2000). Mitochondrial control of cell death. Nat. Med. 6, 513–19CrossRefGoogle ScholarPubMed
Kuhl, D. E., Phelps, M. E., Markham, C. H., Metter, J., Riege, W. H. & Winter, J. (1982). Cerebral metabolism and atrophy in Huntington's disease determined by 18FDG and computed tomographic scan. Ann. Neurol., 12, 425–34CrossRefGoogle ScholarPubMed
Leist, M., Single, B., Castoldi, A. F., Kuhnle, S. & Nicotera, P. (1997). Intracellular ATP concentration: a switch between apoptosis and necrosis. J. Exp. Med., 185, 1481–6CrossRefGoogle ScholarPubMed
Li, L. Y., Luo, X. & Wang, X. (2001). Endonuclease G is an apoptotic DNase when released from mitochondria. Nature, 412, 95–9CrossRefGoogle ScholarPubMed
Liu, X., Kim, C. N., Yang, J., Jemmerson, R. & Wang, X. (1996). Induction of apoptotic program in cell-free extracts: requirement for dATP and cytochrome c. Cell, 86, 147–57CrossRefGoogle ScholarPubMed
Loschen, G., Azzi, A., Richter, C.et al. (1974). Superoxide radicals as precursors of mitochondrial hydrogen peroxide. FEBS Lett., 42, 68–72CrossRefGoogle ScholarPubMed
Luo, X., Budihardjo, I., Zou, H., Slaughter, C. & Wang, X. (1998). Bid, a Bcl2 interacting protein, mediates cytochrome c release from mitochondria in response to activation of cell surface death receptors. Cell, 94, 481–90CrossRefGoogle ScholarPubMed
Mata, M., Fink, D. J., Gainer, H.et al. (1980). Activity-dependent energy metabolism in rat posterior pituitary primarily reflects sodium pump activity. J. Neurochem., 34, 213–15CrossRefGoogle ScholarPubMed
Matthews, R. T., Yang, L., Browne, S., Baik, M. & Beal, M. F. (1998a). Coenzyme Q10 administration increases brain mitochondrial ATP concentrations and exerts neuroprotective effects. Proc. Natl Acad. Sci., USA, 95, 8892–7CrossRefGoogle Scholar
Matthews, R. T., Yang, L., Jenkins, B. G.et al. (1998b). Neuroprotective effects of creatine and cyclocreatine in animal models of Huntington's disease. J. Neurosci., 18, 156–63CrossRefGoogle Scholar
Mazziotta, J. C., Phelps, M. E. & Pahl, J. I. (1987). Reduced cerebral glucose metabolism in asymptomatic subjects at risk for Huntington's disease. N. Engl. J. Med., 316, 356–62CrossRefGoogle ScholarPubMed
Murphy, A. N., Fiskum, G. & Beal, M. F. (1999). Mitochondria in neurodegeneration: bioenergetic function in cell life and death. J. Cereb. Blood Flow Metab., 19, 231–45CrossRefGoogle ScholarPubMed
Narita, M., Shimizu, S., Ito, T.et al. (1998). Bax interacts with the permeability transition pore to induce permeability transition and cytochrome c release in isolated mitochondria. Proc. Natl Acad. Sci., USA, 95, 14681–6CrossRefGoogle ScholarPubMed
Nicklas, W. J., Vyas, I. & Heikkila, R. E. (1985). Inhibition of NADH-linked oxidation in brain mitochondria by 1-methyl-4-phenyl-pyridine, a metabolite of the neurotoxin, 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine. Life Sci., 36, 2503–8CrossRefGoogle Scholar
Orth, M. & Schapira, A. H. V. (2001). Mitochondria and degenerative disorders. Am. J. Med. Genet. (Semin. Med. Genet.), 106, 27–36CrossRefGoogle ScholarPubMed
Panov, A. V., Gutekunst, C- A., Leavitt, B. R.et al. (2002). Early mitochondrial calcium defects in Huntington's disease are a direct effect of polyglutamines. Nat. Neurosci., 5, 731–6CrossRefGoogle ScholarPubMed
Parker, W. D., Boyson, S. J. & Parks, J. K. (1989). Abnormalities of the electron transport chain in idiopathic Parkinson's disease. Ann. Neurol., 26, 719–23CrossRefGoogle ScholarPubMed
Peng, T. I. & Greenamyre, J. T. (1998). Privileged access to mitochondria of calcium influx through N-methyl-D-aspartate receptors. Mol. Pharmacol., 53, 974–80Google ScholarPubMed
Pfeiffer, D. R., Gunter, T. E., Eliseev, R.et al. (2001). Release of Ca2+ from mitochondria via the saturable mechanisms and the permeability transition. IUBMB Life, 52, 205–12CrossRefGoogle ScholarPubMed
Polyak, K., Xin, Y., Zweier, J. L., Kinzler, K. W. & Vogelstein, B. (1997). A model for p53-induced apoptosis. Nature, 289, 300–4CrossRefGoogle Scholar
Pulsinelli, W. A. & Cooper, A. J. L. (1989). Metabolic encephalopathies and coma. In Basic Neurochemistry, ed. G. Siegel, B. Agranoff, R. W. Albers & P. Molinoff, pp. 765–81. New York: Raven Press
Puthalakath, H., Huang, D. C., O'Reilly, L. A., King, S. M. & Strasser, A. (1999). The proapoptotic activity of the Bcl-2 family member Bim is regulated by interaction with the dynein motor complex. Mol. Cel., 3, 287–96CrossRefGoogle ScholarPubMed
Raha, S. & Robinson, B. H. (2001). Mitochondria, oxygen free radicals, and apoptosis. Am. J. Med. Genet. (Semin. Med. Genet.), 106, 62–70CrossRefGoogle ScholarPubMed
Ramsay, R. R., Krueger, M. J. & Youngster, S. K. (1991). Interaction of 1-methyl-4-phenylpyridinium ion (MPP+) and its analogs with the rotenone/piericidin binding site of NADH dehydrogenase. J. Neurochem., 56, 1184–90CrossRefGoogle ScholarPubMed
Ravagnan, L., Roumier, T. & Kroemer, G. (2002). Mitochondria, the killer organells and their weapons. J. Cell Physiol., 192, 131–7CrossRefGoogle Scholar
Schapira, A. H. V., Cooper, J. M., Dexter, D., Jenner, P., Clark, J. B. & Marsden, C. D. (1989). Mitochondrial complex I deficiency in Parkinson's disease. Lancet, i, 1269CrossRefGoogle Scholar
Schapira, A. H. V., Mann, V. M. & Cooper, J. M. (1990). Anatomic and disease specificity of NADH CoQ1 reductase (complex I) deficiency in Parkinson's disease. J. Neurochem., 55, 2142–5CrossRefGoogle ScholarPubMed
Schulz, J. B., Henshaw, D. R., Siwek, D.et al. (1995) Involvement of free radicals in excitotoxicity in vivo. J. Neurochem., 63, 2239–47Google Scholar
Sheehan, J. P., Swerdlow, R. H., Miller, S. W.et al. (1997a). Calcium homeostasis and reactive oxygen species production in cells transformed by mitochondria from individuals with sporadic Alzheimer's disease. J. Neurosci., 17, 4612–22CrossRefGoogle Scholar
Sheehan, J. P., Swerdlow, R. H., Parker, W. D., Miller, S. W., Davis, R. E. & Tuttle, J. B. (1997b). Altered calcium homeostasis in cells transformed by mitochondria from individuals with Parkinson's disease. J. Neurochem., 68, 1221–33CrossRefGoogle Scholar
Sherer, T. B., Betarbet, R., Stout, A. K.et al. (2002). An in vitro model of Parkinson's disease: linking mitochondrial impairment to altered alpha-synuclein metabolism and oxidative damage. J. Neurosci., 22, 7006–15CrossRefGoogle Scholar
Shimizu, S., Narita, M . & Tsujimoto, Y. (1999). Bcl-2 family proteins regulate the release of apoptogenic cytochrome c by the mitochondrial channel VDAC. Nature, 399, 4833–7Google ScholarPubMed
Slater, E. C. (1973). The mechanism of action of the respiratory inhibitor, antimycin. Biochem. Biophys. Act., 301, 129–54Google ScholarPubMed
Stryer, L. (1988). In Biochemistry, pp. 349–426. New York: W. H. Freeman and Company
Stout, A. K., Raphael, H. M., Kanterewicz, B. I.et al. (1998). Glutamate-induced neuron death requires mitochondrial calcium uptake. Nat. Neurosci., 1, 366–73CrossRefGoogle ScholarPubMed
Swerdlow, R. H., Parks, J. K., Cassarino, D. S.et al. (1997). Cybrids in Alzheimer's disease: a cellular model of the disease?Neurology, 49, 918–25CrossRefGoogle ScholarPubMed
Swerdlow, R. H., Parks, J. K., Davis, J. N.et al. (1998). Matrilineal inheritance of complex I dysfunction I in a multigenerational Parkinson's disease family. Ann. Neurol., 44, 873–81CrossRefGoogle Scholar
Tymianski, M., Charlton, M. P., Carlen, P. L. & Tator, C. H. (1993). Source specificity of early calcium neurotoxicity in cultured embryonic spinal neurons. J. Neurosci., 13, 2085–104CrossRefGoogle ScholarPubMed
Vander Heiden, M. G., Chandel, N. S., Schumacker, P. T. & Thompson, C. B. (1999). Bcl-xl prevents cell death following growth factor withdrawal by facilitating mitochondrial ATP/ADP exchange. Mol. Cel., 3, 159–67CrossRefGoogle 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
Volbracht, C., Leist, M. & Nicotera, P. (1999). ATP controls neuronal apoptosis triggered by microtubule breakdown or potassium deprivation. Mol. Med. 5, 477–89Google ScholarPubMed
Vogelstein, B., Lane, D. & Levine, A. J. (2000). Surfing the p53 network. Nature, 408, 307–10CrossRefGoogle ScholarPubMed
Wieloch, T. (2001). Mitochondrial involvement in acute neurodegeneration. IUBMB Life, 52, 247–54CrossRefGoogle ScholarPubMed
Woodfield, K., Ruck, A., Brdiczka, D. & Halestrap, A. P. (1998). Direct demonstration of a specific interaction between cyclophilin-D and the adenine nucleotide translocase confirms their role in the mitochondrial permeability transition. Biochem. J. 336, 287–90CrossRefGoogle ScholarPubMed
Yan, L. J., Levine, R. L. & Sohal, R. S. (1997). Oxidative damage during aging targets mitochondrial aconitase. Proc. Natl Acad. Sci., USA, 94, 11168–72CrossRefGoogle ScholarPubMed
Yue, T. L., Gou, J. L., Lysko, P. G., Cheng, H. Y., Barone, F. C. & Feuerstein, G. (1992). Neuroprotective effects of phenyl-t-butyl-nitrone in getbil global brain ischemia and in cultured rat cerebellar neurons. Brain Res., 574, 193–7CrossRefGoogle ScholarPubMed
Zamzami, N., Susin, S. A., Marchetti, P.et al. (1996). Mitochondrial control of nuclear apoptosis. J. Exp. Med., 183, 1661–72CrossRefGoogle ScholarPubMed
Zhang, Y., Marcillat, O., Giulivi, C., Ernster, L. & Davies, K. J. (1990). The oxidative inactivation of mitochondrial electron transport chain components and ATPase. J. Biol. Chem., 265, 16330–6Google 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.

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.

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.

Available formats
×