Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-26T15:23:52.669Z Has data issue: false hasContentIssue false

8 - Genetically engineered animals

Published online by Cambridge University Press:  04 November 2009

Carolina M. Maier
Affiliation:
Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA
Lilly Hsieh
Affiliation:
Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA
Pak H. Chan
Affiliation:
Department of Neurosurgery and Neurosciences Stanford University Medical Center Palo Alto, CA 93304 USA
Turgut Tatlisumak
Affiliation:
Helsinki University Central Hospital
Marc Fisher
Affiliation:
University of Massachusetts Medical School
Get access

Summary

Introduction

Animal models that recreate specific pathogenic events and their corresponding behavioral outcomes are indispensable tools for exploring the underlying pathophysiologic mechanisms of disease and for investigating therapeutic strategies prior to testing them in human patients. Although rodents have a long tradition as models for human neurological diseases, they have received increasing attention in light of genetic engineering methods that have made it possible to create precisely defined genetic changes. The development of transgenic technology, a tool that allows sophisticated manipulation of the genome, has provided an unprecedented opportunity to expand our understanding of many aspects of neuronal development, function, and disease.

Transgenic animals are specific variants of species following the introduction and/or integration of a new gene or genes into the genome of the host animal. Transgenic technology is now routinely used to increase the level of (overexpress) particular proteins or enzymes in animals or to mutate or inactivate a particular gene using a “knockout” approach. One of the most commonly used animals in transgenic techniques is the mouse, a species in which transgene microinjections into the pronuclei of fertilized oocytes and the subsequent expression of the transgene in the animal have been carefully worked out.

Since their development in the 1980s, the transgenic and knockout technologies have allowed us to examine the regulation of gene expression and the pathophysiology of its alterations.

Type
Chapter
Information
Handbook of Experimental Neurology
Methods and Techniques in Animal Research
, pp. 114 - 131
Publisher: Cambridge University Press
Print publication year: 2006

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

Garcia, AD, Doan, NB, Imura, T, et al. GFAP-expressing progenitors are the principal source of constitutive neurogenesis in adult mouse forebrain. Nature Neurosci. 2004, 7: 1233–1241.CrossRefGoogle ScholarPubMed
Marchis, S, Temoney, F, Erdelvi, F, et al. GABAergic phenotypic differentiation of a subpopulation of subventricular derived migrating progenitors. Eur. J. Neurosci. 2004, 20: 1307–1317.CrossRefGoogle ScholarPubMed
Wen, PH, Shao, X, Shao, Z, et al. Overexpression of wild type but not an FAD mutant presenilin-1 promotes neurogenesis in the hippocampus of adult mice. Neurobiol. Dis. 2002, 10: 8–19.CrossRefGoogle Scholar
Menuet, D, Viengchareun, S, Penfornis, P, et al. Targeted oncogenesis reveals a distinct tissue-specific utilization of alternative promoters of the human mineralocorticoid receptor gene in transgenic mice. J. Biol. Chem. 2000, 275: 7878–7886.CrossRefGoogle ScholarPubMed
Steinbach, JP, Kozmik, Z, Pfeffer, P, et al. Overexpression of Pax5 is not sufficient for neoplastic transformation of mouse neuroectoderm. Int. J. Cancer 2001, 93: 459–467.CrossRefGoogle Scholar
Tong, WM, Ohgaki, H, Huang, H, et al. Null mutation of DNA strand break-binding molecule poly(ADP-ribose) polymerase causes medulloblastomas in p53(− / −) mice. Am. J. Pathol. 2003, 162: 343–352.CrossRefGoogle ScholarPubMed
Moy, LY, Tsai, LH. Cyclin-dependent kinase 5 phosphorylates serine 31 of tyrosine hydroxylase and regulates its stability. J. Biol. Chem. 2004, 279: 54487–54493.CrossRefGoogle ScholarPubMed
Hyun, DH, Lee, M, Hattori, N, et al. Effect of wild-type or mutant Parkin on oxidative damage, nitric oxide, antioxidant defenses, and the proteasome. J. Biol. Chem. 2002, 277: 28572–28577.CrossRefGoogle ScholarPubMed
Saito, A, Hayashi, T, Okuno, S, et al. Modulation of the Omi/HtrA2 signaling pathway after transient focal cerebral ischemia in mouse brains that overexpress SOD1. Brain Res. Mol. Brain Res. 2004, 127: 89–95.CrossRefGoogle ScholarPubMed
Pizoli, CE, Jinnah, HA, Billingsley, ML, et al. Abnormal cerebellar signaling induces dystonia in mice. J. Neurosci. 2002, 22: 7825–7833.CrossRefGoogle ScholarPubMed
Winder, DG, Mansuy, IM, Osman, M, et al. Genetic and pharmacological evidence for a novel, intermediate phase of long-term potentiation suppressed by calcineurin. Cell 1998, 92: 25–37.CrossRefGoogle ScholarPubMed
Casas, C, Sergeant, N, Itier, JM, et al. Massive CA1/2 neuronal loss with intraneuronal and N-terminal truncated Abeta42 accumulation in a novel Alzheimer transgenic model. Am. J. Pathol. 2004, 165: 1289–1300.CrossRefGoogle Scholar
Manley, GT, Binder, DK, Papadopoulos, MC, et al. New insights into water transport and edema in the central nervous system from phenotype analysis of aquaporin-4 null mice. Neuroscience 2004, 129: 983–991.CrossRefGoogle ScholarPubMed
Okamura, N, Suemotu, T, Shiomitsu, T, et al. A novel imaging probe for in vivo detection of neuritic and diffuse amyloid plaques in the brain. J. Mol. Neurosci. 2004, 24: 247–255.CrossRefGoogle Scholar
Richards, JG, Higgins, GA, Ouagazzal, AM, et al. PS2APP transgenic mice, coexpressing hPS2mut and hAPPswe, show age-related cognitive deficits associated with discrete brain amyloid deposition and inflammation. J. Neurosci. 2003, 23: 8989–9003.CrossRefGoogle ScholarPubMed
Menalled, LB, Chesselet, MF. Mouse models of Huntington's disease. Trends Pharmacol. Sci. 2002, 23: 32–39.CrossRefGoogle ScholarPubMed
Janus, C, Chishti, MA, Westaway, D. Transgenic mouse models of Alzheimer's disease. Biochim. Biophys. Acta 2000, 1502: 63–75.CrossRefGoogle ScholarPubMed
Dawson, TM, Dawson, VL. Neuroprotective and neurorestorative strategies for Parkinson's disease. Nature Neurosci. 2002, 5 (Suppl.): 1058–1061.CrossRefGoogle ScholarPubMed
Lim, KL, Dawson, VL, Dawson, TM. The genetics of Parkinson's disease. Curr. Neurol. Neurosci. Rep. 2002, 2: 439–446.CrossRefGoogle ScholarPubMed
Bruijn, LI, Houseweart, MK, Kato, S, et al. Aggregation and motor neuron toxicity of an ALS-linked SOD1 mutant independent from wild-type SOD1. Science 1998, 281: 1851–1854.CrossRefGoogle ScholarPubMed
Rakhit, R, Cunningham, P, Furtos-Matei, A, et al. Oxidation-induced misfolding and aggregation of superoxide dismutase and its implications for amyotrophic lateral sclerosis. J. Biol. Chem. 2002, 277: 47551–47556.CrossRefGoogle ScholarPubMed
Burright, EN, Orr, HT, Clark, HB. Mouse models of human CAG repeat disorders. Brain Pathol. 1997, 7: 965–977.CrossRefGoogle ScholarPubMed
Glaze, DG. Rett syndrome: of girls and mice – lessons for regression in autism. Ment. Retard. Devel. Disabil. Res. Rev. 2004, 10: 154–158.CrossRefGoogle ScholarPubMed
Reeves, RH, Irving, NG, Moran, TH, et al. A mouse model for Down syndrome exhibits learning and behaviour deficits. Nature Genet. 1995, 11: 177–184.CrossRefGoogle ScholarPubMed
Heinzer, AK, McGuinness, MC, Lu, JF, et al. Mouse models and genetic modifiers in X-linked adrenoleukodystrophy. Adv. Exp. Med. Biol. 2003, 544: 75–93.CrossRefGoogle ScholarPubMed
Games, D. Adams, R, Alessandrini, R, et al. Alzheimer-type neuropathology in transgenic mice overexpressing V717 F beta-amyloid precursor protein. Nature 1995, 373: 523–527.CrossRefGoogle Scholar
Hsiao, K, Chapman, P, Nilsen, S, et al. Correlative memory deficits, Abeta elevation, and amyloid plaques in transgenic mice. Science 1996, 274: 99–102.CrossRefGoogle ScholarPubMed
Phinney, AL, Horne, P, Yang, J, et al. Mouse models of Alzheimer's disease: the long and filamentous road. Neurol. Res. 2003, 25: 590–600.CrossRefGoogle Scholar
Ahlijanian, MK, Barrezueta, NX, Williams, RD, et al. Hyperphosphorylated tau and neurofilament and cytoskeletal disruptions in mice overexpressing human p25, an activator of cdk5. Proc. Natl Acad. Sci. USA 2000, 97: 2910–2915.CrossRefGoogle ScholarPubMed
Lewis, J, Dickson, DW, Lin, WL, et al. Enhanced neurofibrillary degeneration in transgenic mice expressing mutant tau and APP. Science 2001, 293: 1487–1491.CrossRefGoogle ScholarPubMed
Holcomb, L, Gordon, MN, McGowan, E, et al. Accelerated Alzheimer-type phenotype in transgenic mice carrying both mutant amyloid precursor protein and presenilin 1 transgenes. Nature Med. 1998, 4: 97–100.CrossRefGoogle ScholarPubMed
Holtzman, DM, Bales, KR, Wu, S, et al. Expression of human apolipoprotein E reduces amyloid-beta deposition in a mouse model of Alzheimer's disease. J. Clin. Invest. 1999, 103: R15–R21.CrossRefGoogle Scholar
Mucke, L, Masliah, E, Yu, GQ, et al. High-level neuronal expression of abeta 1–42 in wild-type human amyloid protein precursor transgenic mice: synaptotoxicity without plaque formation. J. Neurosci. 2000, 20: 4050–4058.CrossRefGoogle ScholarPubMed
Lee, MK, Stirling, W, Xu, Y, et al. Human alpha-synuclein-harboring familial Parkinson's disease-linked Ala-53 → Thr mutation causes neurodegenerative disease with alpha-synuclein aggregation in transgenic mice. Proc. Natl Acad. Sci. USA 2002, 99: 8968–8973.CrossRefGoogle ScholarPubMed
Dauer, W, Kholodilov, N, Vila, M, et al. Resistance of alpha-synuclein null mice to the parkinsonian neurotoxin MPTP. Proc. Natl Acad. Sci. USA 2002, 99: 14524–14529.CrossRefGoogle ScholarPubMed
Barnes, GT, Duyao, MP, Ambrose, CM, et al. Mouse Huntington's disease gene homolog (Hdh). Somat. Cell Mol. Genet. 1994, 20: 87–97.CrossRefGoogle Scholar
Lin, CH, Tallaksen-Greene, S, Chien, WM, et al. Neurological abnormalities in a knock-in mouse model of Huntington's disease. Hum. Mol. Genet. 2001, 10: 137–144.CrossRefGoogle Scholar
Lione, , Carten, RJ, Hunt, MJ, et al. Selective discrimination learning impairments in mice expressing the human Huntington's disease mutation. J. Neurosci. 1999, 19: 10428–10437.CrossRefGoogle ScholarPubMed
Morton, AJ, Lagan, MA, Skepper, JN, et al. Progressive formation of inclusions in the striatum and hippocampus of mice transgenic for the human Huntington's disease mutation. J. Neurocytol. 2000, 29: 679–702.CrossRefGoogle ScholarPubMed
Yamamoto, A, Lucas, JJ, Hen, R. Reversal of neuropathology and motor dysfunction in a conditional model of Huntington's disease. Cell 2000, 101: 57–66.CrossRefGoogle Scholar
Wilson, JM, Khabazian, I, Pow, DV, et al. Decrease in glial glutamate transporter variants and excitatory amino acid receptor down-regulation in a murine model of ALS-PDC. Neuromol. Med. 2003, 3: 105–118.CrossRefGoogle Scholar
Dal Canto, MC, Gurney, ME. Neuropathological changes in two lines of mice carrying a transgene for mutant human Cu, Zn SOD, and in mice overexpressing wild type human SOD: a model of familial amyotrophic lateral sclerosis (FALS). Brain Res. 1995, 676: 25–40.CrossRefGoogle Scholar
Morikawa, E, Mori, H, Kiyama, Y, et al. Attenuation of focal ischemic brain injury in mice deficient in the epsilon1 (NR2A) subunit of NMDA receptor. J. Neurosci. 1998, 18: 9727–9732.CrossRefGoogle ScholarPubMed
Sans, N, Vissel, B, Petralia, RS, et al. Aberrant formation of glutamate receptor complexes in hippocampal neurons of mice lacking the GluR2 AMPA receptor subunit. J. Neurosci. 2003, 23: 9367–9373.CrossRefGoogle ScholarPubMed
Namura, S, Maeno, H, Takami, S, et al. Inhibition of glial glutamate transporter GLT-1 augments brain edema after transient focal cerebral ischemia in mice. Neurosci. Lett. 2002, 324: 117–120.CrossRefGoogle ScholarPubMed
Yang, G, Chan, PH, Chen, J, et al. Human copper-zinc superoxide dismutase transgenic mice are highly resistant to reperfusion injury after focal cerebral ischemia. Stroke 1994, 25: 165–170.CrossRefGoogle ScholarPubMed
Murakami, K, Kondo, T, Kawase, M, et al. Mitochondrial susceptibility to oxidative stress exacerbates cerebral infarction that follows permanent focal cerebral ischemia in mutant mice with manganese superoxide dismutase deficiency. J. Neurosci. 1998, 18: 205–213.CrossRefGoogle ScholarPubMed
Sheng, H, Bart, RD, Oury, TD, et al. Mice overexpressing extracellular superoxide dismutase have increased resistance to focal cerebral ischemia. Neuroscience 1999, 88: 185–191.CrossRefGoogle ScholarPubMed
Samdani, AF, Dawson, TM, Dawson, VL. Nitric oxide synthase in models of focal ischemia. Stroke 1997, 28: 1283–1288.CrossRefGoogle ScholarPubMed
Huang, Z, Huang, PL, Ma, J, et al. Enlarged infarcts in endothelial nitric oxide synthase knockout mice are attenuated by nitro-L-arginine. J. Cereb. Blood Flow Metab. 1996, 16: 981–987.CrossRefGoogle ScholarPubMed
Fannjiang, Y, Kim, CH, Huganir, RL, et al. BAK alters neuronal excitability and can switch from anti- to pro-death function during postnatal development. Devel. Cell 2003, 4: 575–585.CrossRefGoogle ScholarPubMed
Trapp, T, Korhonen, L, Besselmann, M, et al. Transgenic mice overexpressing XIAP in neurons show better outcome after transient cerebral ischemia. Mol. Cell Neurosci. 2003, 23: 302–313.CrossRefGoogle ScholarPubMed
Boutin, H, LeFeuvre, RA, Horai, R, et al. Role of IL-1alpha and IL-1beta in ischemic brain damage. J. Neurosci. 2001, 21: 5528–5534.CrossRefGoogle ScholarPubMed
Friedlander, RM, Gagliardini, V, Hara, H, et al. Expression of a dominant negative mutant of interleukin-1 beta converting enzyme in transgenic mice prevents neuronal cell death induced by trophic factor withdrawal and ischemic brain injury. J. Exp. Med. 1997, 185: 933–940.CrossRefGoogle ScholarPubMed
Soriano, SG, Coxon, A, Wang, YF, et al. Mice deficient in Mac-1 (CD11b/CD18) are less susceptible to cerebral ischemia/reperfusion injury. Stroke 1999, 30: 134–139.CrossRefGoogle ScholarPubMed
Ishikawa, M, Cooper, D, Russell, J, et al. Molecular determinants of the prothrombogenic and inflammatory phenotype assumed by the postischemic cerebral microcirculation. Stroke 2003, 34: 1777–1782.CrossRefGoogle ScholarPubMed
Wang, YF, Tsirka, SE, Strickland, S, et al. Tissue plasminogen activator (tPA) increases neuronal damage after focal cerebral ischemia in wild-type and tPA-deficient mice. Nature Med. 1998, 4: 228–231.CrossRefGoogle ScholarPubMed
Asahi, M, Sumii, T, Fini, ME, et al. Matrix metalloproteinase 2 gene knockout has no effect on acute brain injury after focal ischemia. NeuroReport 2001, 12: 3003–3007.CrossRefGoogle ScholarPubMed
Jo, K, Rutten, B, Bunn, RC, et al. Actinin-associated LIM protein-deficient mice maintain normal development and structure of skeletal muscle. Mol. Cell Biol. 2001, 21: 1682–1687.CrossRefGoogle ScholarPubMed
Dennis, CL, Tinsley, JM, Deconinck, AE, et al. Molecular and functional analysis of the utrophin promoter. Nucleic Acids Res. 1996, 24: 1646–1652.CrossRefGoogle ScholarPubMed
Kanadia, RN, Johnstone, KA, Mankodi, A, et al. A muscleblind knockout model for myotonic dystrophy. Science 2003, 302: 1978–1980.CrossRefGoogle ScholarPubMed
Chan, PH. Reactive oxygen radicals in signaling and damage in the ischemic brain. J. Cereb. Blood Flow Metab. 2001, 21: 2–14.CrossRefGoogle ScholarPubMed
Chan, PH. Epstein, CJ, Li, Y, et al. Transgenic mice and knockout mutants in the study of oxidative stress in brain injury. J. Neurotrauma 1995, 12: 815–824.CrossRefGoogle Scholar
Hesselager, G, Holland, EC. Using mice to decipher the molecular genetics of brain tumors. Neurosurgery 2003, 53: 685–694; discussion 695.CrossRefGoogle ScholarPubMed
Kalaria, RN, Viitanen, M, Kalimo, H, et al. The pathogenesis of CADASIL: an update. J. Neurol. Sci. 2004, 226: 35–39.CrossRefGoogle ScholarPubMed
Longhi, L, Saatman, KE, Raghupathi, R, et al. A review and rationale for the use of genetically engineered animals in the study of traumatic brain injury. J. Cereb. Blood Flow Metab. 2001, 21: 1241–1258.CrossRefGoogle Scholar
McCullough, L, Wu, L, Haughey, N, et al. Neuroprotective function of the PGE2 EP2 receptor in cerebral ischemia. J. Neurosci. 2004, 24: 257–268.CrossRefGoogle ScholarPubMed
Gordon, JW, Scangos, GA, Plotkin, DJ, et al. Genetic transformation of mouse embryos by microinjection of purified DNA. Proc. Natl Acad. Sci. USA 1980, 77: 7380–7384.CrossRefGoogle ScholarPubMed
Wagner, EF. EMBO medal review: On transferring genes into stem cells and mice. EMBO J. 1990, 9: 3024–3032.Google ScholarPubMed
Evans, MJ, Kaufman, MH. Establishment in culture of pluripotential cells from mouse embryos. Nature 1981, 292: 154–156.CrossRefGoogle ScholarPubMed
Humpherys, D, Eggan, K, Akutsu, H, et al. Epigenetic instability in ES cells and cloned mice. Science 2001, 293: 95–97.CrossRefGoogle ScholarPubMed
Elroy-Stein, O, Bernstein, Y, Groner, Y. Overproduction of human Cu/Zn-superoxide dismutase in transfected cells: extenuation of paraquat-mediated cytotoxicity and enhancement of lipid peroxidation. EMBO J. 1986, 5: 615–622.Google ScholarPubMed
Epstein, CJ, Avraham, KB, Lovett, M, et al. Transgenic mice with increased Cu/Zn-superoxide dismutase activity: animal model of dosage effects in Down syndrome. Proc. Natl Acad. Sci. USA 1987, 84: 8044–8048.CrossRefGoogle ScholarPubMed
White, BA. PCR Protocols: Current Methods and Applications. Totowa, NJ: Humana Press, 1993.CrossRefGoogle Scholar
Thomas, KR, Capecchi, MR. Site-directed mutagenesis by gene targeting in mouse embryo-derived stem cells. Cell 1987, 51: 503–512.CrossRefGoogle ScholarPubMed
Li, Y, Huang, TT, Carlson, EJ, et al. Dilated cardiomyopathy and neonatal lethality in mutant mice lacking manganese superoxide dismutase. Nature Genet. 1995, 11: 376–381.CrossRefGoogle ScholarPubMed
Li, Y, Carlson, EJ, Murakami, K, et al. Targeted expression of human CuZn superoxide dismutase gene in mouse central nervous system. J. Neurosci. Methods 1999, 89: 49–55.CrossRefGoogle ScholarPubMed
Chan, PH, Kawase, M, Murakami, K, et al. Overexpression of SOD1 in transgenic rats protects vulnerable neurons against ischemic damage after global cerebral ischemia and reperfusion. J. Neurosci. 1998, 18: 8292–8299.CrossRefGoogle ScholarPubMed
Sung, YH, Song, J, Lee, HW. Functional genomics approach using mice. J. Biochem. Mol. Biol. 2004, 37: 122–132.Google ScholarPubMed
Lee, D, Threadgill, DW. Investigating gene function using mouse models. Curr. Opin. Genet. Devel. 2004, 14: 246–252.CrossRefGoogle ScholarPubMed
Kwan, KM. Conditional alleles in mice: practical considerations for tissue-specific knockouts. Genesis 2002, 32: 49–62.CrossRefGoogle ScholarPubMed
Schnutgen, F, Doerflinger, N, Calleja, C, et al. A directional strategy for monitoring Cre-mediated recombination at the cellular level in the mouse. Nature Biotechnol. 2003, 21: 562–565.CrossRefGoogle ScholarPubMed
Gossen, M, Freundlieb, S, Bender, G, et al. Transcriptional activation by tetracyclines in mammalian cells. Science 1995, 268: 1766–1769.CrossRefGoogle ScholarPubMed
Orban, PC, Chui, D, Marth, JD. Tissue- and site-specific DNA recombination in transgenic mice. Proc. Natl Acad. Sci. USA 1992, 89: 6861–6865.CrossRefGoogle ScholarPubMed
Dymecki, SM. Flp recombinase promotes site-specific DNA recombination in embryonic stem cells and transgenic mice. Proc. Natl Acad. Sci. USA 1996, 93: 6191–6196.CrossRefGoogle ScholarPubMed
Rajewsky, K, Gu, H, Kuhn, R, et al. Conditional gene targeting. J. Clin. Invest. 1996, 98: 600–603.CrossRefGoogle ScholarPubMed
Knuesel, I, Elliott, A, Chen, HJ, et al. A role for synGAP in regulating neuronal apoptosis. Eur. J. Neurosci. 2005, 21: 611–621.CrossRefGoogle ScholarPubMed
Akassoglou, K, Malester, B, Xu, J, et al. Brain-specific deletion of neuropathy target esterase/swisscheese results in neurodegeneration. Proc. Natl Acad. Sci. USA 2004, 101: 5075–5080.CrossRefGoogle ScholarPubMed
Branda, CS, Dymecki, SM. Talking about a revolution: the impact of site-specific recombinases on genetic analyses in mice. Devel. Cell 2004, 6: 7–28.CrossRefGoogle ScholarPubMed
Maier, CM, Chan, PH. Role of superoxide dismutases in oxidative damage and neurodegenerative disorders. Neuroscientist 2002, 8: 323–334.CrossRefGoogle ScholarPubMed
Rosen, DR, Siddique, T, Patterson, D, et al. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 1993, 362: 59–62.CrossRefGoogle ScholarPubMed
Brown, RH Jr. Superoxide dismutase and familial amyotrophic lateral sclerosis: new insights into mechanisms and treatments. Ann. Neurol. 1996, 39: 145–146.CrossRefGoogle ScholarPubMed
Rosen, DR. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 1993, 364: 362.CrossRefGoogle ScholarPubMed
Gurney, ME, Pu, H, Chiu, AY, et al. Motor neuron degeneration in mice that express a human Cu, Zn superoxide dismutase mutation. Science 1994, 264: 1772–1775.CrossRefGoogle ScholarPubMed
Williamson, TL, Cleveland, DW. Slowing of axonal transport is a very early event in the toxicity of ALS-linked SOD1 mutants to motor neurons. Nature Neurosci. 1999, 2: 50–56.CrossRefGoogle ScholarPubMed
Al-Chalabi, A, Leigh, PN. Recent advances in amyotrophic lateral sclerosis. Curr. Opin. Neurol. 2000, 13: 397–405.CrossRefGoogle ScholarPubMed
Iadecola, C, Zhang, F, Niwa, K, et al. SOD1 rescues cerebral endothelial dysfunction in mice overexpressing amyloid precursor protein. Nature Neurosci. 1999, 2: 157–161.CrossRefGoogle ScholarPubMed
Nakao, N, Frodl, EM, Widner, H, et al. Overexpressing Cu/Zn superoxide dismutase enhances survival of transplanted neurons in a rat model of Parkinson's disease. Nature Med. 1995, 1: 226–231.CrossRefGoogle Scholar
Kondo, T, Reaume, AG, Huang, TT, et al. Reduction of CuZn-superoxide dismutase activity exacerbates neuronal cell injury and edema formation after transient focal cerebral ischemia. J. Neurosci. 1997, 17: 4180–4189.CrossRefGoogle ScholarPubMed
Lewen, A, Matz, P, Chan, PH. Free radical pathways in CNS injury. J. Neurotrauma 2000, 17: 871–890.CrossRefGoogle ScholarPubMed
Pineda, JA, Aono, M, Sheng, H, et al. Extracellular superoxide dismutase overexpression improves behavioral outcome from closed head injury in the mouse. J. Neurotrauma 2001, 18: 625–634.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.

  • Genetically engineered animals
    • By Carolina M. Maier, Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA, Lilly Hsieh, Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA, Pak H. Chan, Department of Neurosurgery and Neurosciences Stanford University Medical Center Palo Alto, CA 93304 USA
  • Edited by Turgut Tatlisumak, Marc Fisher
  • Book: Handbook of Experimental Neurology
  • Online publication: 04 November 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541742.008
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.

  • Genetically engineered animals
    • By Carolina M. Maier, Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA, Lilly Hsieh, Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA, Pak H. Chan, Department of Neurosurgery and Neurosciences Stanford University Medical Center Palo Alto, CA 93304 USA
  • Edited by Turgut Tatlisumak, Marc Fisher
  • Book: Handbook of Experimental Neurology
  • Online publication: 04 November 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541742.008
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.

  • Genetically engineered animals
    • By Carolina M. Maier, Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA, Lilly Hsieh, Department of Neurosurgery and Neurological Sciences Stanford University Medical School 1201 Welch Rd MSLS P357 Stanford, CA 94305 USA, Pak H. Chan, Department of Neurosurgery and Neurosciences Stanford University Medical Center Palo Alto, CA 93304 USA
  • Edited by Turgut Tatlisumak, Marc Fisher
  • Book: Handbook of Experimental Neurology
  • Online publication: 04 November 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511541742.008
Available formats
×