+ All Categories
Home > Documents > TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and...

TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and...

Date post: 04-Jul-2019
Category:
Upload: nguyenthu
View: 215 times
Download: 0 times
Share this document with a friend
108
Mutant Cu,Zn Superoxide Dismutase in Amyotrophic Lateral Sclerosis Molecular Mechanisms of Neurotoxicity Doctoral dissertation To be presented by permission of the Faculty of Natural and Environmental Sciences of the University of Kuopio for public examination in Auditorium, Mediteknia building, University of Kuopio, on Saturday 3 rd May 2008, at 12 noon Department of Neurobiology A.I. Virtanen Institute for Molecular Sciences University of Kuopio TONI AHTONIEMI JOKA KUOPIO 2008 KUOPION YLIOPISTON JULKAISUJA G. - A.I. VIRTANEN -INSTITUUTTI 61 KUOPIO UNIVERSITY PUBLICATIONS G. A.I. VIRTANEN INSTITUTE FOR MOLECULAR SCIENCES 61
Transcript
Page 1: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

Mutant Cu,Zn Superoxide Dismutase inAmyotrophic Lateral Sclerosis

Molecular Mechanisms of Neurotoxicity

Doctoral dissertation

To be presented by permission of the Faculty of Natural and Environmental

Sciences of the University of Kuopio for public examination in

Auditorium, Mediteknia building, University of Kuopio,

on Saturday 3rd May 2008, at 12 noon

Department of NeurobiologyA.I. Virtanen Institute for Molecular Sciences

University of Kuopio

TONI AHTONIEMI

JOKAKUOPIO 2008

KUOPION YLIOPISTON JULKAISUJA G. - A.I. VIRTANEN -INSTITUUTTI 61KUOPIO UNIVERSITY PUBLICATIONS G.

A.I. VIRTANEN INSTITUTE FOR MOLECULAR SCIENCES 61

Page 2: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

Distributor : Kuopio University Library P.O. Box 1627 FI-70211 KUOPIO FINLAND Tel. +358 17 163 430 Fax +358 17 163 410 http://www.uku.fi/kirjasto/julkaisutoiminta/julkmyyn.html

Series Editors: Research Director Olli Gröhn, Ph.D. Department of Neurobiology A.I . Virtanen Institute for Molecular Sciences

Research Director Michael Courtney, Ph.D. Department of Neurobiology A.I . Virtanen Institute for Molecular Sciences

Author’s address: Department of Neurobiology A.I . Virtanen Institute for Molecular Sciences University of Kuopio P.O. Box 1627 FI-70211 KUOPIO FINLAND

Supervisors: Professor Jari Koistinaho, M.D., Ph.D. Department of Neurobiology A.I . Virtanen Institute for Molecular Sciences University of Kuopio

Docent Gundars Goldsteins, Ph.D. Department of Neurobiology A.I . Virtanen Institute for Molecular Sciences University of Kuopio

Reviewers: Docent Pekka Rauhala, M.D., Ph.D. Institute of Biomedicine University of Helsinki Helsinki, Finland

Professor Dan Linholm, M.D., Ph.D. Minerva Research Institute University of Helsinki Helsinki, Finland

Opponent: Professor Michael Thomas Heneka, M.D. Department of Neurology University of Münster Münster, Germany

ISBN 978-951-27-1120-8ISBN 978-951-27-1102-4 (PDF)ISSN 1458-7335

KopijyväKuopio 2008Finland

Page 3: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

Ahtoniemi, Toni. Mutant Cu,Zn Superoxide Dismutase in Amyotrophic Lateral Sclerosis:Molecular Mechanisms of Neurotoxicity. Kuopio University Publications G. - A.I. VirtanenInstitute for Molecular Sciences 61. 200 . 107 p.ISBN 978-951-27-1120-8ISBN 978-951-27-1102-4 (PDF)ISSN 1458-7335

ABSTRACTAmyotrophic lateral sclerosis (ALS) is an untreatable and fatal neurodegenerative disease thatleads to muscle atrophy, paralysis of voluntary muscles and death. ALS is charactherized bythe selective destruction of upper and lower motor neurons in the spinal cord, brain stem andmotor cortex. The majority of ALS cases are sporadic, whereas 5–10% of cases have a geneticcomponent and are familial. The typical age of onset for both forms is between 50 and 60years and prevalence is approximately 4-6 in 100 000 individuals per year. The causes formost cases of ALS are unknown and the clinical course is highly variable, suggesting thatmultiple factors underlie the disease mechanism.

Mutations in the ubiquitously expressed protein, Cu,Zn-superoxide dismutase (SOD1),are associated with about 20% of familial ALS cases. Because the pathology and clinicalsymptoms of familial and sporadic ALS cannot be distinguished, transgenic animal modelsover-expressing mutant SOD1 offer a valuable tool for understanding the pathogenicmechanisms shared by both sporadic and familial forms of ALS. Importantly, severalpathogenic SOD1 mutations do not affect SOD1 activity significantly, and a hypothesis for 'atoxic gain of function' of the mutated protein rather than a lack of its antioxidant function, hasbeen postulated.

In this work, we used transgenic ALS rats and mice to address the role of SOD1 inALS pathogenesis by analyzing the proteomic profile in the spinal cord and by analyzingstability and oxidation state of human mutant SOD1 through the disease progression. Inaddition, we set out to investigate the role of mutant SOD1 in mitochondria.

Results showed that mutant SOD1 is oxidized and destabilized in the affected tissuesof the transgenic animals. A role of aggregation was further supported by the finding ofincreased chaperone expression in proteome profiling. Moreover, SOD1 may have adeleterious role in the intermembrane space of mitochondria, however, not because ofaggregation, but caused instead by uncontrolled activity of the enzyme that can lead toincreased production of harmful reactive oxygen species and damage to mitochondria. Inaddition, we also tested an anti-oxidant/inflammatory drug treatment with pyrrolidinedithiocarbamate for SOD1 transgenic ALS rats. However, the result of this drug treatmentwas unexpected as the treatment did not provide neuroprotection, but in opposite to thehypothesis, accelerated the disease progression. The mechanism of action showed a newtarget for the drug as the beneficial anti-oxidative and anti-inflammatory effects wereoverridden by a harmful inhibition of immunoproteasome induction.

The results of this thesis show the importance of mammalian specific proteasomecomponent in the disease pathogenesis of ALS. Moreover, our results demonstrate a newmitochondrial target and mechanism for SOD1 neurotoxicity that applies both to sporadic andfamilial ALS cases.

National Library of Medicine Classification: WL 359, WE 550, QU 58.5, QU 140, QU 350,QZ 180Medical Subject Headings: Neurodegenerative Diseases/etiology; Motor Neuron Disease;Amyotrophic Lateral Sclerosis; Superoxide Dismutase; Mutant Proteins; Proteomics; SpinalCord; Mitochondria; Oxidative Stress; Reactive Oxygen Species; Molecular Chaperones;Proteasome Endopeptidase Complex; Disease Models, Animal; Mice; Rats; Drug Therapy;Antioxidants; Pyrrolidines

8

Page 4: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor
Page 5: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

"Nothing shocks me. I'm a scientist."Indiana Jones

Page 6: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor
Page 7: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

ACKNOWLEDGEMENTS

This study was carried out in the Molecular Brain Research Group, Department ofNeurobiology, A. I. Virtanen Institute for Molecular Sciences, University of Kuopio, duringthe years 2002-2007.

I wish to express my sincere gratitude to my principal supervisor, Professor Jari Koistinaho,M.D., Ph.D., for his supervision, wonderful insight on neurobiology and excellent leadership.It has truly been a priviledge to work in the MBR-group for these past years. I couldn’t evenbegin to imagine a better choice for PhD-studies than the MBR-group with Jari as asupervisor. I also wish to thank my supervisor Docent Gundars Goldsteins, Ph.D., for hissupervision, unbelievable technical and scientific insight and for scrutinizing andtroubleshooting. Gundars, Gunu or Mr. Wizard, your input on my views of science has beentruly remarkable and your views and advices will be always heard on every word.

I am grateful to the official reviewers of this thesis, Professor Dan Lindholm, M.D., Ph.D., atMinerva Institute for Medical Research, University of Helsinki and Docent Pekka Rauhala,M.D., Ph.D. at Institute of Biomedicine, University of Helsinki for their valuable evaluationof this thesis.

I am thankful to my co-authors for their contiribution that is beyond any price. Velta Keksa-Goldsteine, M.Sc., and Merja Jaronen, M.Sc., thank you for your work in the protein lab -keep Gunu's lab a happy and a productive place. I want to thank from the bottom of my heartKatja Kanninen, M.Sc., and Tarja Malm, Ph.D., who have shared many memorable momentsthrough these years. Katja and Tarja, your contribution and friendship goes beyond any scaleimaginable. I am also deeply thankful to the co-authoring past members of the A.I.VirtanenInstitute, Eija Seppälä, Ph.D., Egils Arens, Ph.D., and Karl Åkerman, Ph.D., as well as to co-authors outside AIVI, Antero Salminen, Ph.D., at the Department of Neurology, SeppoAuriola, Ph.D., at the Department of Pharmaceutical Chemistry, University of Kuopio,Caterina Bendotti, Ph.D., Department of Neuroscience, Istituto di Ricerche FarmacologicheMario Negri, Milan, Italy, and Pak Chan, Ph.D., Department of Neurosurgery, StanfordUniversity School of Medicine, Stanford, California, USA.

My warm thanks goes to our lovely technicians Mirka Tikkanen and Laila Kaskela for theirhelp, contribution and friendship in and outside the lab. I would also like to thank all themembers of the MBR-group; Susanna Boman, VMD, Riikka Heikkinen, M.Sc., Georges FulKuh, M.Sc., Johanna Magga, Ph.D., Anu Muona, Ph.D., Rea Pihlaja, M.Sc., Eveliina Pollari,M.Sc., Yuriu Pomeshchick, M.D., and Taisia Rolova, M.Sc. Many thanks goes to the pastmembers of MBR-group and participants of the legendary New Orleans SFN tour of 2003;Antti “Possum Jenkins” Nurmi, Ph.D., and Suvi Leskinen, M.Sc. The previous generations ofMBR-PhDs Tiina “Tikka” Kauppinen, Ph.D., Kaisa Kurkinen, Ph.D. and Nina Vartiainen,Ph.D. are also cordially acknowledged. A special thank you goes to a past member of theMBR-group and my current boss Docent Milla Koistinaho Ph.D. for giving me a job-opportunity of a lifetime at Medeia. From AIVI staff I would also like to thank Docent RiittaKeinänen and Mrs. Sari Koskelo for their help in countless affairs and in their dedication toMBR-group.

Love and gratitude goes for my parents Tuija and Hannu, who have always encouraged me tostudy hard and reach as high as possible. I would like to thank my brother Janne for hisfriendship, views on life and for being an excellent brother for the years past and to come. Iwould also like to thank Jannes’s lovely wife Maiju, even lovelier daughter Ella and

Page 8: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

newcomer Akseli. Lots of love and kisses to my littlesister Ellinoora. Families Nygård andJärvelä are also warmly acknowledged.

Finally, all my love goes to my wife Pauliina and sons Tomas and Leevi. Thank you for yourlove, support and generally just for putting up with me during the finalizing steps of thisthesis... In the end family is what matters the most and you are the counterbalance in my life.

Kuopio, April 2008

Toni Ahtoniemi

Page 9: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

ABBREVIATIONS

ALS amyotrophic lateral sclerosisANG angiogeninBSA bovine serum albuminCNS central nervous systemCOX-2 cyclooxygenase-2COX4 cytochrome c oxidase 4 subunitDCF dichlorodihydrofluorescein diacetateDEAE diethylamino ethanolDNPH dinitrophenylhydrazineDTT dithiothreitolEAAT2 excitatory amino acid transporter 2ECL enhanced chemiluminescenceEMSA electrophoretic mobility shift assayFTD frontotemporal dementiaGEF guanine exchange factorGFAP glial fibrillary acidic proteinGLT-1 glutamate transporter 1GSK glycogen synthase kinaseHPR horse radish peroxidaseHSP heat shock proteinIGF-1 insulin like growth factor 1IL-1 interleukin 1ICV intracerebroventicularmalPEG mono-Methyl polyethylene glycol 5'000 2-maleimidoethyl etherMND motor neuron diseaseNF- B nuclear factor BNO nitric oxidePBP progressive bulbar palsyPDI protein disulphide isomerasePDTC pyrrolidine dithiocarbamatePLS primary lateral sclerosisPMA progressive muscular atrophyPMSF phenylmethylsulphonyl fluoridePVDF polyvinylidene fluorideROS reactive oxygen speciesSDS-PAGE sodium dodecyl sulphate polyacrylamide gel electrophoresisSMA spinal muscular atrophySOD superoxide dismutaseTG transgenicTNF tumor necrosis factorVEGF vascular endothelial growth factorWT wild type

Page 10: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor
Page 11: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

LIST OF ORIGINAL PUBLICATIONS

This thesis is based on the following original publications, which in text are referred to by

their roman numerals.

I Ahtoniemi T, Goldsteins G, Keksa-Goldsteine V, Malm T, Kanninen K, Salminen A,

and Koistinaho J. Pyrrolidine dithiocarbamate inhibits induction of

immunoproteasome and decreases survival in a rat model of amyotrophic lateral

sclerosis. Mol Pharmacol. 2007 Jan;71(1):30-37

II Ahtoniemi T, Jaronen M, Keksa-Goldsteine V, Goldsteins G, and Koistinaho J.

(2008). Mutant SOD1 from spinal cord of G93A rats binds to inner mitochondrial

membrane and increases ROS production. Submitted.

III Goldsteins G, Keksa-Goldsteine V*, Ahtoniemi T*, Jaronen M, Egils A, Åkerman K,

Chan PH, and Koistinaho J. Deleterious role of superoxide dismutase in the

mitochondrial intermembrane space. J. Biol. Chem. 2008 Mar 28;283(13):8446-52

IV Ahtoniemi T, Seppälä E, Goldsteins G, Auriola S, Bendotti C, and Koistinaho J.

Proteomic analysis of protein expression and oxidation in a mouse model of ALS.

Manuscript.

*Contributed equally to the work

Page 12: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor
Page 13: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

TABLE OF CONTENTS

1. INTRODUCTION ............................................................................................................ 152. REVIEW OF THE LITERATURE ................................................................................... 17

2.1 Amyotrophic lateral sclerosis (ALS) ........................................................................... 172.1.1 ALS and other motor neuron diseases................................................................... 17

2.1.1.1 Motor neuron system ......................................................................... 172.1.1.2 Motor neuron diseases ....................................................................... 18

2.1.2 Characteristics of ALS ......................................................................................... 192.1.3 Genetics of ALS ................................................................................................... 21

2.1.3.1 Sporadic and Familial ALS ................................................................ 212.1.3.2 ALS1 - SOD1 .................................................................................... 232.1.3.3 ALS2, ALS4 and ALS5 - Juvenile forms ALS ................................... 252.1.3.4 ALS3, ALS6 and ALS7 with classical late-onset phenotype ............... 262.1.3.5 ALS with dementia ............................................................................ 262.1.3.6 ALS8 and progressive lower motor neuron disease - atypical ALS ..... 272.1.3.7 VEGF and ANG ................................................................................ 27

2.1.4 Models of ALS..................................................................................................... 282.1.4.1 Transgenic SOD1 models................................................................... 282.1.4.2 Other in vivo models of motor neuron degeneration ........................... 302.1.4.3 In vitro models of ALS ...................................................................... 32

2.2 Mechanisms for motor neuron cell death ..................................................................... 332.2.1 Oxidative damage ................................................................................................ 33

2.2.1.1 SOD1 activity .................................................................................... 332.2.1.2 Aberrant SOD1 activity ..................................................................... 34

2.2.2 Protein Aggregation ............................................................................................. 362.2.2.1 Aggregates ......................................................................................... 362.2.2.2 Proteasome and Immunoproteasome .................................................. 372.2.2.3 Chaperones ........................................................................................ 382.2.2.7 Neurofilaments and axonal transport .................................................. 39

2.2.3 Glutamate excitotoxicity ...................................................................................... 392.2.4 Inflammation........................................................................................................ 402.2.5 Mitochondria........................................................................................................ 412.2.6 Role of non-neuronal cells .................................................................................... 432.2.7 Pathway of motor neuron cell death in ALS ......................................................... 45

2.3 Therapeutics for ALS .................................................................................................. 472.3.1 Drug treatments .................................................................................................... 482.3.2 Growth factors ..................................................................................................... 492.3.3 Gene therapies...................................................................................................... 502.3.4 Stem cell therapies ............................................................................................... 502.3.5 Pyrrolidine dithiocarbamate (PDTC) .................................................................... 52

3. AIMS OF THE STUDY ................................................................................................... 544. MATERIALS AND METHODS ...................................................................................... 55

4.1 Animals (I-IV) ............................................................................................................ 554.2 PDTC treatment (I) ..................................................................................................... 564.3 Mitochondria (II,III) ................................................................................................... 57

4.3.1 Isolation of mitochondria ..................................................................................... 574.3.2 Functional integry of the isolated mitochondria .................................................... 584.3.3 Isolation of mitoplasts .......................................................................................... 584.3.4 Exposure of mitoplasts with cytosolic homogenates of G93A-SOD1 rat tissues.... 584.3.5 Measurement of ROS production ......................................................................... 594.3.6 Isolation of intermembrane space and measurement of SOD1 activity .................. 59

Page 14: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

4.3.7 SOD activity with zymography ............................................................................ 594.4 Western blotting (I-III)................................................................................................ 60

4.4.1 Sample preparation .............................................................................................. 604.4.2 Electrophoresis and transfer ................................................................................. 604.4.3 malPEG modification of free cysteines ................................................................. 604.4.4 Antibodies ............................................................................................................ 61

4.5 Immunohistochemistry (I) ........................................................................................... 614.6 Electrophoretic mobility shift assay (I) ........................................................................ 624.7 Atomic absorption spectrophotometry (I) .................................................................... 624.8 Proteasomal activity (I) ............................................................................................... 624.9 Proteomics (IV) .......................................................................................................... 63

4.9.1 Protein extraction ................................................................................................. 634.9.2 Carbonyl derivatization and detection................................................................... 634.9.3 Two dimensional electrophoresis ......................................................................... 644.9.4 In-gel digestion .................................................................................................... 644.9.5 Mass spectrometry ............................................................................................... 65

4.10 Flow cytometry (III) ................................................................................................. 654.10.1 Peroxide production in mouse blood lymphocytes .............................................. 654.10.2 Antimycin A-induced apoptosis in lymphocytes ................................................. 65

4.11 Isolation and purification of human SOD1 (II,III) ..................................................... 664.12 Measurements of cytochrome c catalysed peroxidation (III) ...................................... 664.13 Statistical analysis (I-IV) ........................................................................................... 66

5. RESULTS ........................................................................................................................ 675.1 PDTC reduced survival of G93A-SOD1 ALS rats without affecting NF- B (I) ........... 675.2 Copper levels were increased in ALS rat tissues and were further increased in spinalcords by PDTC treatment (I) ............................................................................................. 685.3 PDTC inhibited immunoproteasome (I) ...................................................................... 685.4 PDTC upregulated GLT-1 (I) ...................................................................................... 695.5 PDTC prevented glial immunoproteasome induction (I) .............................................. 695.6 Mutant SOD1 was oxidized and destabilized in spinal cords of G93A-SOD1 rodentmodels (II,IV) ................................................................................................................... 705.7 PDI was upregulated and its levels inversely correlated with the levels of cysteinereduced SOD1 (II,IV) ....................................................................................................... 715.8 Mitochondrial SOD1 levels were the highest in the spinal cord of G93A rats (II) ........ 725.9 Mutant SOD1 bound to mitoplasts and enhanced ROS production (II) ........................ 725.10 Mutant SOD1 increased hydrogen peroxide production in the intermembrane space ofmitochondria (III) ............................................................................................................. 735.11 Mutant SOD1 activity and ROS production were increased in spinal cordmitochondria of G93A-SOD1 mice (III) ........................................................................... 75

6. DISCUSSION .................................................................................................................. 776.1 PDTC inhibits immunoproteasome induction resulting in reduced survival of ALS rats(I) ..................................................................................................................................... 776.2 Mutant SOD1 oxidation and destabilization precede aggregation and loss of activity (II,IV) .................................................................................................................................... 796.3 Destabilized mutant SOD1 associates with inner membrane of mitochondria andincreases ROS production (II) ........................................................................................... 816.4 Elevated SOD1 activity in the intermembrane space leads to increased hydrogenperoxide production resulting in cytochome c catalyzed oxidation (III) ............................. 826.5 Hypothesized role of destabilized mutant SOD1 toxicity in mitochondria ................... 85

7. SUMMARY AND CONCLUSIONS ................................................................................ 878. REFERENCES ................................................................................................................. 89

Page 15: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

15

1. INTRODUCTION

Amyotrophic lateral sclerosis (ALS) is a neurodegenerative disease characterized by loss of

upper and lower motor neurons in the spinal cord, brain stem and motor cortex. Loss of motor

neurons leads to muscle atrophy, paralysis of voluntary muscles and death in 3-5 years from

onset of the disease with failure of the respiratory muscles usually being the fatal event

(Kandel et al., 1991). What makes ALS a very cruel disease is the fact that although the body

becomes completely paralysed and the ability to communicate is lost, the mind and senses

remain intact as sensory neurons or autonomic neurons are not affected (Kandel et al., 1991).

5–10% of ALS cases are inherited (familial) whereas the majority of cases have no genetic

component (sporadic) (Kurland and Mulder, 1955). The typical age of onset for familial form

is 46 years and 56 years for sporadic with prevalence of approximately 4-6 in 100 000

individuals per year (Camu et al., 1999; Yoshida et al., 1986). This corresponds to

approximately 150 new affected individuals in Finland every year, whereas the total number

of all affected people is approximately 350 in Finland, 5000 in the United Kingdom and 30

000 in the United States or Europe. The causes for most cases of ALS are unknown and the

clinical course is highly variable, suggesting that multiple factors underlie the disease

mechanism.

Mutations in the ubiquitously expressed protein, Cu,Zn superoxide dismutase (SOD1)

are associated with about 20% of familial ALS cases (Rosen et al., 1993). Because the

pathology and clinical symptoms of familial and sporadic ALS cannot be distinguished,

transgenic animal models over-expressing mutant SOD1 (Bruijn et al., 1997b; Gurney et al.,

1994; Howland et al., 2002; Jonsson et al., 2004; Nagai et al., 2007; Ripps et al., 1995; Wang

et al., 2003; Wang et al., 2005; Wong et al., 1995) offer a valuable tool for understanding the

pathogenic mechanisms shared by both sporadic and familial forms of ALS. Importantly,

several pathogenic SOD1 mutations do not affect SOD1 activity significantly, and a

hypothesis for 'a toxic gain of function' of the mutated protein rather than a lack of its

antioxidant function, has been postulated (Gurney et al., 1994; Wong et al., 1995). The nature

of this gained toxic function is not known, even though several putative pathogenic

mechanisms have been discovered, including formation of protein aggregates, saturation of

proteasome and protein folding chaperones, mislocalization and aggregation of

neurofilaments, inflammation, increased radical generation and oxidative damage,

mitochondrial dysfunction, and pro-apoptotic alterations (Boillee et al., 2006a). Despite the

many proposed disease mechanisms, the underlying initiative cause of the mutant SOD1

toxicity is still unknown. Mutations of SOD1 may affect the stability of the enzyme making it

Page 16: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

16

more prone to aggregate or mutations may alter the catalytical site allowing aberrant

substrates to enter the catalytical site and cause unwanted oxidative reactions (Cleveland and

Rothstein, 2001). In recent studies, mitochondrial localisation of mutant SOD1 has been

implicated in ALS pathogenesis (Bergemalm et al., 2006; Deng et al., 2006; Ferri et al., 2006;

Liu et al., 2004; Vijayvergiya et al., 2005) and increased recruitment of mutant SOD1 to

mitochondria in the spinal cord might be the basis of the specific cell death of motor neurons.

However, the exact mechanisms of the selectivity and toxicity are not entirely clear.

This study was carried out to further address the role of SOD1 in ALS pathogenesis by

analyzing the proteomic profile in the spinal cord of ALS mice and by analyzing the stability

and oxidation state of human mutant SOD1 through the disease progression in transgenic

ALS mice and rats. In addition, we set out to investigate the deleterious role of mutant SOD1

in mitochondria. Finally, as oxidative damage, inflammation and apoptosis play major roles in

the pathology of ALS, we tested whether anti-inflammatory and anti-oxidative drug treatment

with pyrrolidine dithiocarbamate might have neuroprotective effects on G93A-SOD1

transgenic rats.

Page 17: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

17

2. REVIEW OF THE LITERATURE

2.1 Amyotrophic lateral sclerosis (ALS)

2.1.1 ALS and other motor neuron diseases

2.1.1.1 Motor neuron system

The motor neuron system gives us the means to interact with the world that surrounds us from

the very basic aspects such as breathing, chewing and moving to more complex motor

performances such as controlling the air flow and the shape of oral cavity to form sound as in

speech or singing, or to delicately control movement of finger and hand to be used in writing

or playing an instrument. The neural components of the motor system extend from the highest

reaches of the cerebral cortex as upper motor neurons of the motor cortex to innervate spinal

cord and to farthest terminals of the motor axons as lower motor neurons of spinal cord

connect to muscles. There are two types of lower motor neurons: alpha motor neurons and

gamma motor neurons. Alpha motor neurons directly trigger muscle contraction and gamma

motor neurons innervate and activate intrafusal muscle fibers, which provide information and

additional control on the force of contraction and length of the muscle (Squire et al., 2003a).

Each lower alpha motor neuron descending along the spinal cord or brain stem sends

an axon to one muscle to innervate muscle fibers (from a few to a hundred or more) forming a

motor unit. As there are two types of voluntary muscle fibers - red for aerobic long-lasting

performances and white for anaerobic explosive and fast movements, there are also different

subtypes of motor units: slow fatigue resistant motor units, fast fatiguable and fast fatigue

resistant motor units. Motor neurons of the fast units are generally bigger and have a larger

diameter, faster conducting axons and generate high frequency bursts of action potentials.

Slow units have a smaller diameter and more slowly conducting axons with relatively steady

and lower frequency activity (Squire et al., 2003a).

Motor neurons are activated by interneurons of different motor programs descending

from the forebrain and the brain stem allowing activation of motor neurons with great

precision. Interneurons and motor neurons form neuronal networks containing information for

specific motor performances such as swallowing, walking or breathing and as the

corresponding network is activated, by will or reflex, the given function of the network is

executed. These networks contain motor programs for various activities form lying

horizontally in bed, which requires little activity to more complex programs like walking,

which requires sequentially activated motorneurons/muscles to act in concert with sensory

receptors for keeping the balance. The spinal cord contains motor programs for locomotion

Page 18: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

18

and for protective reflexes, the brain stem has the programs for swallowing, chewing,

breathing and fast saccadic eye movement, whereas motor programs for fine motor skills such

as speech and control of hands and fingers are located in the motor cortex (Squire et al.,

2003b).

2.1.1.2 Motor neuron diseases

Motor neuron diseases (MND) are a class of diseases that affect the motor system leading to

muscle atrophy and in the worst case scenario to paralysis of the muscles as the motor

neurons that give the signals to muscles are lost. Motor neuron diseases can be divided into

three different categories on the basis of which class of motor neurons (upper motor neurons,

lower motor neurons or both) the disease affects. Spinal muscular atrophy (SMA) is a pure

lower motor neuron disorder; progressive bulbar palsy (PBP) and progressive muscular

atrophy (PMA) have isolated lower motor neuron signs. Primary lateral sclerosis (PLS) has

only upper motor neuron involvement, whereas amyotrophic lateral sclerosis (ALS) affects

both upper and lower motor neurons and this is one of the key characteristics of ALS

(Donaghy, 1999; Rocha et al., 2005).

ALS is the most common form of MNDs with the prevalence of 4-6 affected

individuals per 100 000 (Yoshida et al., 1986) and also the most devastating, as ALS usually

leads to complete paralysis and death in 3-5 years from the diagnosis of the disease and there

is no known cure. What makes ALS so devastating is the fact that it only affects the motor

neurons while there is no cognitive decline nor is the autonomic system affected, leaving the

person affected by ALS capable to follow the progression of the disease and the deteriation of

the body and losing the ability speak and to swallow while the mind still remains as sane as it

is possible in that situation (Kandel et al., 1991).

Other types of progressive adult MNDs include primary lateral sclerosis (PLS),

progressive muscular atrophy (PMA), progressive bulbar palsy (PBP) and only lower motor

neurons involving spinal muscular atrophy (SMA). PLS affects only upper motor neurons

leading to spasticity starting in the legs and ascending to the arms and finally to the bulbar

muscles. PLS has an average age of onset at 50 years and slow disease progression of more

than 15 years. PLS is not fatal, but affects the quality of life and PLS may often develop into

full scale ALS (Pringle et al., 1992). PBP affects lower motor neurons, which control bulbar

muscles leading to pharyngeal muscle weakness and to almost always evident but less

prominent limb weakness with both lower and upper motor neuron signs. PMA has only

lower motor neuron signs with muscle involvement mainly in limbs but also body trunk and

Page 19: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

19

bulbar region can be affected. Both PBP and PMA often progress to classic ALS (Rocha et

al., 2005). SMA is a pure lower motor neuron disease that can be genetic or sporadic and

generally predominate in the legs. SMA has various forms, with different ages of onset,

patterns of inheritance, and severity and progression of symptoms (Harding and Thomas,

1980).

2.1.2 Characteristics of ALS

ALS was initially characterized first by a french neurologist and physican Jean-Martin

Charcot in 1869 and was known initially as Charcot's sclerosis (Charcot and Joffory, 1869).

The fact that ALS was characterized already in 1869 makes ALS as a described disease some

37 years older than Alzheimer's disease, which was reported to the medical society in 1907 by

Alois Alzheimer (Alzheimer, 1907a; Alzheimer, 1907b). Although ALS as a disease is more

rare than AD and it was still described almost forty years sooner than Alzheimer's disease,

emphasizes the importance of the motor system. On the other hand ALS is also more visible

and leaves more evident marks. What Charcot described already 139 years ago was the

observation of a distinct "myelin pallor" in the lateral portions of the spinal cord, representing

the degeneration and loss of motor neurons as they descend from the brain to connect to the

lower motor neurons within the spinal cord.

Clinical features of ALS include progressive muscle weakness, atrophy and spasticity

resulting in the end in complete paralysis of voluntary muscles as motor neurons degenerate

(Figure 1) (Kandel et al., 1991). Muscle weakness and atrophy is mainly caused by the

degeneration of lower motor neurons whereas spasticity reflects the loss of upper motor

neurons. Respiratory failure caused by the denervation of respiratory muscles or pneumonia is

usually the fatal event. ALS is fatal and there is no known cure. The most common first

symptom is weakness of one arm or leg and clumsiness of the hands. Legs can have cramps.

Muscle weakness and atrophy spread from the distal parts of the limbs to the proximal

direction to other limbs, to muscles taking care of breathing and to the bulbar region.

Alternatively, muscle weakness starts from the bulbar region and spreads in the opposite

direction. The first symptom of the bulbar region is weakening of speech and swallowing. The

disease progresses to atrophy of the voluntary muscles and paralysis without any sensory

symptoms, nor is the autonomic nervous system affected and there is no cognitive decline

(Kandel et al., 1991). However, there is also selectivity between the destruction of motor

neurons as the motor neurons controlling the bladder and sphincters are spared and also eye

Page 20: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

20

movement is relatively spared and is affected only in the very late stage of the disease

(Kandel et al., 1991).

Figure 1. In ALS both upper and lower motor neurons degenerate. Motor neuronslocated in the brain, brain stem, and spinal cord serve as controlling units and vitalcommunication links between the nervous system and the voluntary muscles of the body.Messages from motor neurons in the brain (upper motor neurons) are transmitted to motorneurons in the spinal cord (lower motor neurons) and from them to particular muscles. InALS, both the upper motor neurons and the lower motor neurons degenerate or die, ceasing tosend messages to muscles. Unable to function, the muscles gradually weaken, waste away(atrophy), and twitch (fasciculations). Eventually, the ability of the brain to start and controlvoluntary movement is lost. As bulbar motor neurons are lost speaking, swallowing and facialexpressions are affected. Death is usually caused by respiratory failure. Modified from:http://www.als-mda.org/publications/fa-als.html.

Motor cortex:Upper motor neurons

Brain stem:Lower (Bulbar)motor neurons

nerves, axons

Rib musclesinvolved inbreathing

Spinal cord:Lowermotorneurons

tongue

muscles of arm

muscles of leg

Page 21: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

21

ALS is normally diagnosed by ruling out all other possibilities. The number of ALS cases

diagnosed every year is 1-2 per 100 000 individuals and the prevalence or the proportion of

affected persons in the population is 4-6 per 100 000 (Yoshida et al., 1986). This corresponds

to some 350 affected individuals in Finland, 5000 in the United Kingdom, 30 000 in the

United States or Europe. The Median age of onset for ALS is 55.

2.1.3 Genetics of ALS

2.1.3.1 Sporadic and Familial ALS

Actually ALS is not just one disease, but belongs to a group of heterogeneous disorders that

affect both upper and lower motor neurons (Table 1). The majority of ALS cases,

approximately 90%, are sporadic and appear without any known genetic component. The

remaining 10% of cases are inherited dominantly and are called familial ALS. Despite the

differences in genetics, sporadic and familial ALS are clinically indistinguishable and there

are only minor variations in age at onset, sex ratio, survival and the frequency with which

onset occurred in the lower extremities. Hence the recognition of the familial form usually

depends on diagnosis of the disease in other family members.

Although the differences between familial and sproradic ALS are minor, the following

are very interesting: The mean age of onset in familial cases is ~46 years and is on average 10

years earlier than in cases with sporadic ALS (~56 years) (Camu et al., 1999). The observed

male to female ratio in familial ALS is 1:1, while sporadic ALS has an unexplained male

predominance of 1.5:1 reported worldwide (Haverkamp et al., 1995).

The idea for the division of ALS into sporadic and familial forms is based on the

initial discovery and publication by Kurland and Mulder from 1955 showing that in about

10% of ALS cases there was a family history with Mendelian genetics (Kurland and Mulder,

1955). This division may however be challenged as the understanding for the genes of smaller

effect and environmental factors has improved and ALS cases showing seemingly sporadic

appearance can have genetic component and linkage. In fact, recently mutations were found

from TAR DNA binding protein of both sporadic and familial forms of ALS, though the

function of the protein in the CNS and role in ALS pathogenesis remains unknown

(Sreedharan et al., 2008). Also the emergence of genome wide screens of ALS patients will

most likely bring new insight to the ALS genetics of the sporadic forms as well (Blauw et al.,

2008).

Page 22: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

22

Table 1. Genetics of ALS. The identified chromosomal loci leading to ALS. Modified from(Boillee et al., 2006a).

Disease Locus Gene (Protein/Function) Heredity OnsetSporadic ALSSALS None None identified None AdultTypical ALSALS1

ALS3ALS6ALS7

21q22.1

18q2116q1220p13

SOD1 (Cu,Zn superoxide dismutase). Converts superoxide to water and hydrogen peroxide.UnknownUnknownUnknown

Dominant

DominantDominantDominant

Adult. 20% of inherited cases

AdultAdultAdult

Juvenile ALSALS2

ALS4

ALS5

2q33

9q34

15q15.1- q21.1

ALS2/alsin. Guanine exchange factor for Rab5 and RAc1. Organization of actin cytoskeleton / vesicle trafficking.SETX (Senataxin). Putative DNA/RNA helicase.

Unknown

Recessive

Dominant

Recessive

Juvenile Heterogeneous disease.

Juvenile. (Recessive mutations cause ataxia-oculomotor apraxia type 2Juvenile

ALS with dementiaALS-FTD

ALS-FTDP

9q21-22

17q21.1

Unknown

MAPT (Tau) Microtubule associated protein

Dominant

Dominant

Adult. ALS with frontotemporal dementia (FTD)Adult. ALS-FTD and Parkinson's Disease

Atypical ALSALS8

Progressivelower motorneurondisease

20q13.3

2p13

VABP (VAMP-associated proteinB). May be involved in vesiculartrafficking.

DCTN1 (Dynactin 1). Axonal transport of cellular organelles and proteins.

Dominant

Dominat

Adult. Heterogenous disease. Most cases with tremor; some typical ALS; 25% is late-onset spinal muscular atrophyAdult. Vocal fold paralysis; atrophy of hands and facial muscles

Better understanding of the factors influencing inheritance such as multiple effects of

single genes (pleiotropy), the interaction of multiple genes with each other (epistasis), the

interaction of genes with environmental factors, splice variants of genes, variations in copy

number and post translational modifications (as reviewed in Simpson and Al-Chalabi, 2006),

have shown that the 'one gene, one trait' model of diseases can be regarded as too simple for

Page 23: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

23

complete description of disease risk, though it is useful for identifying genes for Mendelian

genetics. Taking these things into consideration, it is not surprising that the identified

Mendelian genes which cause ALS account for a very small percentage of cases, despite the

use of modern gene mapping methods. High genetic heterogenity and complex interactions

between genetic and environmental factors make ALS a multifactorial complex disease.

A genetic link has been determined for about 10% of ALS cases, but the mechanism

for inheritance is far from unambiguous. The identified chromosomal loci with mutations

causing ALS or ALS-like symptoms have been defined as ALS1-8, as well as for progressive

lower motor neuron disease, ALS with frontotemporal dementia (ALS-FTD) and ALS-FTD

with Parkinson's disease (ALS-FTDP) (as reviewed in Gros-Louis et al., 2006). Out of these

listed loci, six have identified genes with Mendelian genetics, namely: ALS1, ALS2, ALS4,

ALS8, ALS-FTDP and progressive lower motor neuron disease. Also, mutations in

angiogenin (ANG), vascular endothelial growth factor (VEGF) and sequence variants in

neurofilament genes have been reported. As a common factor, some of the identified genes

seem to be involved in intracellular trafficking, axonal transport and RNA metabolism.

However, the nomenclature on the genetics of familial ALS cases can be misleading,

as only ALS1, ALS3, ALS6, ALS7 mutations in ANG and VEGF and some of the ALS8

cases have the classical ALS phenotype with late onset and degeneration of both upper and

lower motor neurons that leads to progressive paralysis. Whereas ALS2, ALS4, ALS5 may

have juvenile onset, ALS8 and progressive lower motor neuron disease have only lower

motor neuron signs and ALS-FTD and ALS-FTDP feature dementia.

2.1.3.2 ALS1 - SOD1

The most common form of inherited ALS, about 20% of familial ALS, is caused by mutations

at chromosome 21q22.1 in the gene encoding protein Cu,Zn superoxide dismutase (SOD1),

also known as ALS1. These mutations correspond to 1-2% of all ALS cases. The function of

this ubiquitously expressed, 153 amino acid residues long cytoplasmic homodimeric enzyme

is to convert superoxide anion, a free radical of reactive oxygen species, to water and

hydrogen peroxide, which is further on detoxified by catalase and glutathione peroxidase

(Fridovich, 1986). The discovery that SOD1 mutations cause ALS was published in Nature in

1993 and it was a major discovery since it was the first gene shown to dominantly cause ALS

(Rosen et al., 1993). However, it was not obvious at all how SOD1, an antioxidative enzyme

expressed in all cell types and tissues throughout the body, could cause the selective

degeneration of motor neurons.

Page 24: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

24

Missense mutations in the SOD1 gene lead to replacement of single amino acid

residues in the protein and, with some exceptions, cause dominant inheritance of the disease.

No mutations have been found in healthy people, except for mutation D90A, in which aspartic

acid in position 90 of SOD1 protein is converted to alanine. When usually SOD1 mutations

cause dominantly inherited disease, the D90A causes ALS in the scandinavian population

only when the mutation is inherited recessively from both parents and expressed

homozygously (Andersen et al., 1995; Sjalander et al., 1995). However, in some populations

with other ethnic backgrounds, the D90A mutation causes dominantly inherited ALS.

Therefore, the D90A mutation is a quite remarkable exception to the rule. It has been

suggested that in scandinavian populations the SOD1 gene with D90A mutation is linked to a

protective gene as the mutation arises from a single founder, however, the gene and

mechanism are not known (Al-Chalabi et al., 1998; Parton et al., 2002).

On the whole, the SOD1 mutations are scattered throughout the primary and three-

dimensional structure of the protein and up to date over 100 mutations have been found

(Valentine et al., 2005). Although soon after the discovery of SOD1 mutations a decrease of

dismutase activity was reported in ALS-patients (Deng et al., 1993; Orrell et al., 1995), the

lack of dismutase activity can not be the primary cause of ALS as some of the mutations do

not affect SOD1's normal enzymatic activity. Therefore it has been hypothesized that

mutations in SOD1 cause the disease through a toxic gain of function (Wong et al., 1995). A

complete list of mutations can be found at an online database for ALS genetics at

http://alsod.iop.kcl.ac.uk/Als/index.aspx. All SOD1 mutations are dominant except for D90A,

which can be either dominant or recessive (Andersen et al., 1996). Different SOD1 mutations

can cause distinct phenotypes differing in age of onset, progression and clinical symptoms.

The A4V mutation is the most common and unfortunately it also gives rise to the most

aggressive form of familial ALS with a mean survival of only one year after onset (Deng et

al., 1993). In contrast, the H46R mutation located within the copper binding domain leads to a

mild form of ALS with an average life expectancy of 18 years after disease onset (Aoki et al.,

1993; Ratovitski et al., 1999). What makes the matter even more perplexing is that mutation

H48Q, which is adjacent to the slow progressing ALS causing H46R, leads to a severe form

of ALS with rapid disease progression (Enayat et al., 1995). Moreover, mutations G37R and

L38V are predicted to have earlier onset different from mutations associated with the

aggressive phenotype, such as A4V (Cudkowicz et al., 1997). Considering the variation of

disease progression among different mutants and the fact that D90A causes ALS either

dominantly or recessively depending on population, it is evident that the clinical phenotype is

Page 25: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

25

modified also by genetic or environmental factors other than SOD1 missense mutations.

Unfortunately, no genetic modifiers have been found (Broom et al., 2006).

Despite the fact that different SOD1 mutations cause considerable variations in disease

phenotype and that SOD1 mutations explain only 1-2% of all ALS cases, ALS research has

heavily focused on ALS1 caused by SOD1 mutations. This is mainly because SOD1

mutations have allowed scientists to develop transgenic animal models expressing mutant

SOD1. Transgenic mice and rats overexpressing mutant SOD1 develop ALS-like symptoms

and can be used as a disease model for ALS (Bruijn et al., 1997b; Gurney et al., 1994;

Howland et al., 2002; Jonsson et al., 2004; Nagai et al., 2001; Ripps et al., 1995; Wang et al.,

2003; Wang et al., 2005; Wong et al., 1995). Most of our knowledge on the pathological

mechanisms of ALS are based on the research done by using these models.

2.1.3.3 ALS2, ALS4 and ALS5 - Juvenile forms ALS

From the discovery of the first ALS causing gene - SOD1, it took scientists nearly ten years to

identify the second gene associated with ALS. The second gene named ALS2, located at

chromosome 2q33, was linked to a rare, recessively inherited and slowly progressing juvenile

form of ALS (Hadano et al., 2001; Hentati et al., 1994). Patients in families of Arabic origin

developed juvenile onset (from 3 to 23 years) of progressive spasticity of the limbs, facial

and pharyngeal muscles, all caused by a mutation in the ALS2 gene. Altogether ten mutations

have been reported for the ALS2 gene and eight out of ten mutations are frameshift mutations,

which lead to premature termination of the transcript and a truncated protein. One nonsense

mutation and one splice variant site mutation have also been reported (as reviewed in Gros-

Louis et al., 2006). This has lead to the conclusion that loss of function of the ALS2 encoded

proteins is causing the disease.

The ALS2 gene spans 80 kbp of human genomic DNA and is predicted to encode a

184 kDa protein, named alsin, consisting of 1657 amino acids. Alsin has multiple motifs

homologous to guanine-nucleotide exchange factors (GEF) and it has been shown to function

as a GEF for Rab5 and RAc1 GTPase through the VPS9 domain linking alsin to the

organization of the actin cytoskeleton and vesicle trafficking (Kunita et al., 2004; Otomo et

al., 2003). A common feature for all found ALS2 mutations is the loss of VPS9-associated

GEF function suggesting that alsin mutations result in a deficit in intracellular trafficking

(Kunita et al., 2004; Otomo et al., 2003). However, the loss of alsin in knockout mice does

not lead to major motor deficits consistent with ALS or other MNDs (Cai et al., 2005).

Interestingly, alsin can also bind to mutant SOD1 and give neuroprotection to motor neuronal

Page 26: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

26

cells against mutant SOD1 toxicity; by studying this interaction the role of these mutations in

ALS pathogenesis may be clarified (Kanekura et al., 2004).

ALS4 is a rare autosomal dominant form of juvenile ALS with linkage to chromosome

9q34 (Chance et al., 1998) where different missense mutations in the senataxin gene (SETX)

were found (Chen et al., 2004). Only heterozygous missense mutations in this gene are linked

to ALS, whereas homozygous deletions including missense, nonsense and deleterious

mutations are associated to an ALS unrelated disorder called ataxia-oculomotor apraxia type 2

(AOA2) (Moreira et al., 2004). As recessive deletion mutations in SETX cause AOA2, then

ALS4 is likely be to caused by gain of function of SETX.

ALS5 linked families show a similar disease phenotype as that of ALS2, except that in

ALS5 there is no spasticity of limb, facial and tongue muscles. Genetic analyses have shown

that type ALS5 is not related to ALS2 at 2q33 but to a chromosome location 15q15.1-q21.1

making ALS5 a distinct genetic entity (Hentati et al., 1998).

2.1.3.4 ALS3, ALS6 and ALS7 with classical late-onset phenotype

Genome wide screens have identified a locus with no relation to SOD1 mutations in

chromosome 18q21 for ALS3 (Hand et al., 2002), 16p12 for ALS6 and 20p13 for ALS7

(Sapp et al., 2003). In contrast to ALS2, which causes a juvenile form of ALS, ALS3, ALS6

and ALS7 give rise to classical ALS with late-onset and progressive paralysis with both upper

and lower motor neuron involvement. In fact, ALS3 was the first reported adult-onset

dominant ALS locus since ALS1. However, the genes with causing mutations have not been

identified.

2.1.3.5 ALS with dementia

ALS with frontotemporal dementia (ALS-FTD) and ALS-FTD with Parkinson's disease

(ALS-FTDP) are cases of motor neuron degeneration that occur in patients with FTD or FTD

and Parkinson's disease. In a set of families in which persons develop both ALS and FTD, a

genetic locus that is linked to ALS with FTD was identified on human chromosome 9q21-q22

(Hosler et al., 2000), whereas families with ALS alone did not link to this locus. In ALS-

FTDP, different mutations in the chromosome 17q21.1 of microtubule associated protein tau

gene (MATP) have been identified (Hutton et al., 1998). Tau has the function of stabilizing

microtubules, promoting their assembly and regulating transport of vesicles and organelles

along the microtubules by binding to tubules and modulating their stability (Rademakers et

al., 2004). However, there is considerable variation in clinical and pathological presentations

Page 27: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

27

of patients and not all patients with ALS-FTDP have MAPT mutations, suggesting genetic

heterogeneity as in sporadic ALS (Hutton et al., 1998).

2.1.3.6 ALS8 and progressive lower motor neuron disease - atypical ALS

A missense P56S mutation in the vesicle-associated membrane protein B gene (VABP) in

chromosome 20q13.33 gives rise to autosomal dominant late-onset atypical ALS8. The

phenotype is characterized by slow progression of the disease and late onset with lower motor

neuron symptoms (Nishimura et al., 2004). Atypical signs are tremor and absence of upper

motor neuron involvement. A few cases have a typical ALS phenotype and some 25% of

cases are late-onset spinal muscular atrophy. VABP encodes a 33 kDa protein VAMP-B,

which is a vesicle membrane protein that can associate with microtubules, suggesting that

mutations in this gene may lead to dysfunction in intracellular membrane trafficking and to

variable MNDs (Nishimura et al., 2004).

Progressive lower motor neuron disease is a rare autosomal dominant form of MND,

where some but not all symptoms overlap with ALS. This form of MND has been linked to

missense mutations of dynactin1 gene (DCTN1) in chromosome 2p13 (Puls et al., 2003). The

DCTN gene encodes dynactin, an axonal transport protein, and missense mutations in the

gene are predicted to distort the folding of dynactin's microtubule binding domain, thus

suggesting that dysfunction of dynactin-mediated transport can lead to motor neuron disease

(Puls et al., 2003).

2.1.3.7 VEGF and ANG

VEGF is a growth factor that promotes the formation of blood vessels and can function also

as a neurotrophic factor: VEGF showed an implication to ALS in an animal model, where

deletion of hypoxia response element of VEGF in mouse resulted in an ALS like phenotype,

possibly through chronic neuronal ischemia and loss of direct neurotrophic effect of VEGF

(Oosthuyse et al., 2001). It is of interest as well that crossbreeding of these mice with G93A-

SOD1 mutant mice accelerates the disease progression, indicating that VEGF may be a

modifier for motor neuron degeneration in SOD1 ALS-mouse. Screening for mutations of

VEGF gene and regions of the promoter in patients showed a 1.8-fold increased risk of

developing ALS in a Belgian, Swedish and a British population (Lambrechts et al., 2003;

Terry et al., 2004), but not in other populations (Brockington et al., 2005; Gros-Louis et al.,

2003).

Page 28: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

28

Mutations in ANG have also been linked to ALS, which together with VEGF,

highlights the role of angiogenesis in motor neuron degeneration. However, missense

mutations of ANG in ALS patients were restricted only to Irish and Scottish populations and

it is still unclear how these mutations affect ANG and provoke MND. Moreover, it is not

known whether ANG has neurotrophic properties (Greenway et al., 2006).

2.1.4 Models of ALS

2.1.4.1 Transgenic SOD1 models

The discovery of SOD1 and the fact that SOD1 mutations cause ALS dominantly made it

possible for scientists to create transgenic mouse model for ALS. To date, overexpression of

either G37R, G85R, G86R, G93A, G127X, H46R/H48Q or H46R/H48Q/H63G/H120G

mutant SOD1 in mice (Bruijn et al., 1997b; Gurney et al., 1994; Howland et al., 2002;

Jonsson et al., 2004; Ripps et al., 1995; Wang et al., 2003; Wang et al., 2005; Wong et al.,

1995) and G93A or H46R mutant SOD1 in rats (Howland et al., 2002; Nagai et al., 2001)

have been shown to cause a neurodegenerative disease similar to human ALS. As the

mutations have distinct phenotypical features in humans, also these models with different

mutations vary in age of onset, disease progression and histopathological features, and thus

reflect human ALS. Moreover, the penetrance in diffrent populations can be mimicked by

varying the mouse strain carrying the mutation (Kunst et al., 2000). On the other hand, the

survival times of these mice vary greatly from 4 months to over a year, depending maybe not

so much on the mutation but rather on the levels of mutant SOD1 expression.

The first paper on transgenic mice expressing human mutant SOD1 with G93A

mutation was published in 1994 (Gurney et al., 1994), just one year after the initial discovery

of SOD1 mutations. The results from that paper were highly significant as it was shown that

first of all, these mice developed ALS-like symptoms and secondly, they developed the

disease despite markedly elevated SOD1 activity and therefore gave the first evidence that

SOD1-linked ALS is not caused by loss of dismutase activity. The second set of transgenic

mice expressed SOD1 with G37R mutation (Wong et al., 1995). These mice developed ALS

like symptoms as well, although G37R-SOD1 retained nearly full enzymatic activity and thus

taken into account the findings made by Gurney et al. 1994 it was concluded that the toxicity

of SOD1 mutants has to arise from the property of a toxic subunit, not from the reduction of

dismutase activity. The loss of function theory was further disproved as SOD1 knock out

mice, in which endogenous murine SOD1 gene was deleted, did not develop motor neuron

Page 29: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

29

disease (Reaume et al., 1996). The final nail for the loss of function theory came from Bruijn

et al. 1998 showing that elimination of SOD1 activity or overexpression of human wild-type

SOD1 in the presence of mutant G85R-SOD1 with reduced dismutase activity in mice does

not protect from the disease and, surprisingly, it can even accelerate it (Deng et al., 2006;

Jaarsma et al., 2000) and thus toxicity is independent of SOD1 activity. In addition, Bruijn et

al. also found protein aggregates containing SOD1 to be a common pathological feature for

SOD1 mutations that otherwise give different phenotypical features (Bruijn et al., 1998). It is

undisputedly clear now that mutations in SOD1 do not cause ALS as a loss of dismutase

activity but through a toxic gain of function. However, the mechanism of how mutant SOD1

exerts toxicity to motor neurons is still uncertain. In addition to many different mutant SOD1

expressing strains, also SOD1 knock out mice (Reaume et al., 1996) and human wild type

SOD1 overexpressing mice (Jaarsma et al., 2000) and rats (Chan et al., 1998) are used greatly

in ALS research as they serve as good controls for the transgene over-expression and for the

role of endogenous SOD1.

SOD1 is a ubiquitously expressed protein and one of the key questions in ALS

research is how only the motor neurons are selectively destroyed while there is no pathology

in other tissues. The fact that SOD1 is ubiquitously expressed raises the possibility that the

toxicity is not coming from the motor neurons themselves but also from the non-neuronal

glial cells surrounding the motor neurons. Although mainly motor neurons are degenerating,

there is also pathology present in astrocytes already in the early phase of the disease (Bruijn et

al., 1997b). To address the role of astocytes, a line of transgenic mice were created that

expressed SOD1 only in astrocytes. These mice had high levels of mutant G86R-SOD1 in

astrocytes driven by a GFAP promoter. Although the mice had increased astrocytosis with

aging they did not develop motor neuron degeneration, thus the authors concluded that

expression of mutant SOD1 in the neurons is critical for the initiation of the disease (Gong et

al., 2000). In another set of transgenic mice the expression of mutant SOD1 was restricted to

neurons alone either by neurofilament promoter (Pramatarova et al., 2001) or by neural

specific enolase promoters (Lino et al., 2002), but the outcome of these trials was that

neuron-restricted expression of mutant SOD1 does not cause pathology or motor neuron

disease. However, some doubts remained as the neuron restricted expression of mutant SOD1

may have resulted in too low protein levels to yield disease. Recently, also neuron specific

expression of human mutant SOD1 was shown to induce motor neuron death in mice

(Jaarsma et al., 2008). Nevertheless, a more definitive answer to the contribution of different

cells to ALS pathogenesis came from a study with chimeric mice that were mixtures of

mutant SOD1 expressing cells and normal wild type cells showing that toxicity to motor

Page 30: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

30

neurons requires damage from mutant SOD1 to surrounding non-neuronal cells (Clement et

al., 2003). In fact, expression of mutant SOD1 in motor neurons at levels which cause early

onset and rapidly progressing disease if expressed ubiquitously, do not cause cell-autonomous

degeneration or death of individual motor neurons.

The most recent development in SOD1 models of ALS has been the creation of mice

carrying a mutant SOD1 gene flanked by loxP, sites allowing the deletion of mutant SOD1

gene by Cre recombinase enzyme (Boillee et al., 2006b). Tissue specific expression of Cre

recombinase either in motor neurons or microglia, and hence the deletion of mutant SOD1 in

respective cells, has shown that mutant SOD1 expression in motor neurons and non-neuronal

neighbors have a different contribution to the disease onset and progression; Mutant SOD1

within microglial cells accelerates disease progression while mutant action within the motor

neurons determines onset and progression of early disease (Boillee et al., 2006b). The role of

mutant SOD1 toxicity within different cells of the CNS is reviewed more extensively in

chapter 2.2.6 Role of non-neuronal cells.

In addition to mouse models, there is also an invertebrate model available for studying

the toxic effect of SOD1 mutations in Caenohabditis elegans, a nematode roundworm.

Although C. elegans is only 1mm long, we must not underestimate the power of C. elegans

models or as Professor of Bioinformatics, Garry Wong from the University of Kuopio put it:"

A worm is not a mouse that is not a man." In fact, aspects of mutant SOD1 toxicity have been

modelled in C. elegans as worms expressing mutant SOD1 showed greater vulnerability to

oxidative stress, and under oxidative stress the mutant forms, but not human wild-type SOD1,

formed potent aggregates to muscles (Oeda et al., 2001). Human mutant or wt SOD1 has also

been expressed in Drosophila (fruit fly) motor neurons, where it however showed no toxic

effect but extended lifespan by 40% (Elia et al., 1999; Parkes et al., 1998).

2.1.4.2 Other in vivo models of motor neuron degeneration

Before the emergence of SOD1 transgenic rodent models, no other model could completely

replicate disease progression as thoroughly as the SOD1 models and as these models failed to

replicate the disease, treatment successes from these models were not carried to human trials

for treatment of ALS. The in vivo rodent models used in ALS research prior to the SOD1

models include axotonomy induced motor-neuron death and some naturally occurring

mutations in mice (as reviewed by Elliott, 1999).

When performed in neonatal animals, direct trauma to the motor nerve axon by

peripheral nerve transsection (axotomy) results in apoptotic cell death of all motor neurons

Page 31: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

31

whose axons were severed. Although axotomy induced motor neuron cell death is invaluable

for studying apoptosis, its relevance to ALS is less certain as the injury caused by axotomy is

acute and the events following may differ a lot from the pathways that are activated in chronic

ALS pathology.

In addition to SOD1 transgenic ALS models also other naturally occurring genetic

rodent models are available for motor neuron degeneration research that have been used for

preclinical testing of agents for human ALS. The wobbler mouse represents a phenotype

characterized by progressive forelimb weakness beginning at about one month of age and

animals survive to the age of one year (Andrews et al., 1974; Mitsumoto and Bradley, 1982).

Pathology includes axonopathy with proximal axonal degeneration as well as neuropathy with

vacuolar changes within anterior horn cells of the spinal cord. However, opposite to ALS, the

pathology is limited only to the spinal cord with limited involvement in brain. The wobbler

phenotype has autosomal recessive inheritance and the gene responsible is Vps54 (vacuolar-

vesicular protein sorting 54) involved in vesicular trafficking (Schmitt-John et al., 2005).

The progressive motor neuronopathy (pmn) mice have a recessively inherited

mutation in tubulin chaperone E gene (Bommel et al., 2002; Martin et al., 2002) and these

mice develop pelvic and hind limb weakness and die by 7 weeks of age (Schmalbruch et al.,

1991). Pathologically the phenotype is characterized by a prominent distal motor neuron

axonopathy. However, the motor neuron soma is relatively spared and in this regard the

pathology is dissimilar to ALS.

The neuromuscular degeneration (nmd) mouse is another autosomal recessive model

of spontaneous progressive motor weakness (Cook et al., 1995). These mice develop rapidly

progressive weakness in their hindlimbs beginning at two weeks of age as motor neurons

degenerate in the lumbar spinal cord. These mice rarely survive past four weeks. The genetic

defect has been identified as a single amino acid deletion and spice donor site mutation in the

gene encoding a ubiquitously expressed ATPase/DNA helicase, also known as SMbp2 (Cox

et al., 1998).

The motor neuron disease mouse, or the mnd mouse, has dominantly inherited

autosomal motor neuron disease with late onset. The mnd mouse is characterized by onset in

the hindlimbs with stiffness, atrophy and paralysis starting at 5-11 months of age and lifespan

of 14 months (Messer and Flaherty, 1986). Pathology indicates neuronal swelling with

cytoplasmic inclusions and motor neuron degeneration in the spinal cord, hypoglossal nuclei

and motor cortex (Messer et al., 1987) but also retinal degeneration (Messer et al., 1993). The

gene is located in chromosome 8 and is a homolog for gene CLN8 encoding a putative

membrane protein with yet unknown functions (Ranta et al., 1999).

Page 32: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

32

Although these mouse models do not mimic all the features of human ALS as

extensively as the SOD1 models, the advantage of these models is that they have generalized

and naturally occurring motor weakness over a more chronic time course with gradual

progression of the disease. However, in comparison to human ALS, the nmd has a different

temporal time course with very early onset, the mnd model has differences in spatial patterns

as also retinal neurons are degenerating and the pmn has quite many dissimilarities in

pathology as only axonopathy is occurring. Differences may suggest that these models are

different disorders from ALS altogether. The wobbler and mnd models exhibit a clinical

course and pathology more closely resembling human motor neuron disease, but it is not

known whether similar molecular or biochemical defects underlie these conditions and human

ALS.

2.1.4.3 In vitro models of ALS

In vitro systems offer ease of manipulation of cells by direct pharmacological administration

or by gene transfections. However, preparation of pure motor neuron cultures is complex as

identification and isolation of motor neurons is difficult. In addition, the neurons used for this

preparation have to be isolated from embryonic or late neonatal time points (Hanson et al.,

1998; Martinou et al., 1992). The lifespan of motor neurons in cultures is also short allowing

better assessment of acute rather than chronic injuries. Pure cultures also do not allow

interaction of motor neurons with other neurons or glia, although, this can also be the whole

point of making pure cultures.

Maybe the best in vitro model of ALS and SOD1 mediated toxicity so far is

microinjection of mutant SOD1 to primary cultured neurons (Durham et al., 1997). The

expression of microinjected mutant SOD1 cDNA results naturally in protein expression, but

also in selective killing of motor neurons but not sensory neurons, whereas cDNA of wild-

type SOD1 does not result in any neurotoxicity. In addition, expression of mutant SOD1

cDNA, but not wild-type SOD1, results in formation of protein aggregates, which is followed

by motor neuron cell death (Bruening et al., 1999; Durham et al., 1997). These findings lead

to the initial proposition that aggregates may have a role in SOD1 mediated toxicity.

Organotypic slices of spinal cord can also be used as a model system for studying

ALS. Slices are prepared from 9 day old mice and motor neurons in these cultures can survive

for up to 3 months (Rothstein et al., 1993). The advantage of the slice model is that some of

the spinal cord structures such as dorsal and ventral horns are preserved and the slice

preserves also some of the neuron-neuron and neuron-glia cell interactions. However,

Page 33: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

33

dissecting the slice from the animal results in multiple axotonomies, leaving motor neurons

deafferented and although many motor neurons survive preparation, it is possible that the

procedure is selective in motor neuron survival (Elliott, 1999).

For the cultures or slices to be used as a model for ALS, the cells have to be

manipulated in order to reflect the disease state. Glutamate excititoxicity is one of the mostly

used methods to cause motor neuron degeneration either by inhibition of glutamate transport

(Rothstein et al., 1993) or by causing excitotoxicity directly by application of N-methyl-D-

aspartic acid (NMDA) or NMDA agonist (Annis and Vaughn, 1998; Delfs et al., 1997).

Agents that block glutamate receptors, inhibit glutamate release or decrease glutamate

synthesis are capable of preventing this form of motor neuron death in vitro (Annis and

Vaughn, 1998; Delfs et al., 1997; Rothstein and Kuncl, 1995).

2.2 Mechanisms for motor neuron cell death

2.2.1 Oxidative damage

2.2.1.1 SOD1 activity

After the landmark discovery that mutations in SOD1 are a cause of ~20% of familial ALS

cases fourteen years ago (Rosen et al., 1993), it was shown that SOD1 activity in patient

blood is reduced (Deng et al., 1993; Orrell et al., 1995) and it was hypothesized that loss of

SOD1 activity, and thus increased levels of superoxide radicals, was central to the disease.

However, soon after it was realized that mutations in SOD1 do not cause ALS through loss of

activity for the following reasons: transgenic mice with mutant SOD1 expression developed

progressive motor neuron disease despite markedly elevated SOD1 activity levels (Gurney et

al., 1994), SOD1 knock out mice in which endogenous SOD1 is completely deleted do not

develop overt motor neuron disease (Reaume et al., 1996), some SOD1 mutants retain full

specific activity (Borchelt et al., 1994) and neither the age of onset or rapidity of disease

progression correlates with dismutase activity levels (Bowling et al., 1995; Cleveland et al.,

1995). The inevitable conclusion is that mutations in SOD1 do not cause ALS through loss of

activity but through toxic gain of function.

Although the loss of activity and increased levels of superoxide anions was not proven

to cause the disease, the oxidative damage might be caused by unwanted oxidative reactions

caused by mutant SOD1. Normally dismutation of superoxide anion is catalysed by SOD1 in

two asymmetric steps by the reactive copper atom, which is alternately reduced and oxidized

by superoxide (figure 2a) (Fridovich, 1986). Even the wild-type SOD1 can exhibit additional

Page 34: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

34

enzymatic activities (Liochev and Fridovich, 2000) and the most obvious hypothesis of how

over 100 different mutations could mediate toxicity is that mutations in SOD1 result in a less

tightly folded protein conformation, allowing greater access of substrates to the reactive

copper and thus the possible catalysis of unwanted oxidative reactions of abnormal substrates.

Some likely abnormal substrates to give oxidative damage include hydrogen peroxide

(H2O2) and peroxynitritre (-ONOO) and the first hypothesized theories on unwanted oxidative

reactions of SOD1 (as reviewed in Cleveland and Rothstein, 2001) were the peroxidase

hypothesis (Wiedau-Pazos et al., 1996) and two peroxynitrate theories (Beckman et al., 1993;

Estevez et al., 1999). The unwanted reactions of these substrates and the dismutase activity of

SOD1 are presented in figure 2.

2.2.1.2 Aberrant SOD1 activity

The first suggested abnormal oxidative reaction of mutant SOD1 included peroxynitrite as a

substrate (Beckman et al., 1993). Peroxynitrite can be formed spontaneously from superoxide

and nitric oxide and if it is used as a substrate by SOD1 it will yield protein tyrosine nitration

(figure 2c). In the case of zinc-depleted mutant SOD1, superoxide anions are formed by

mutant SOD1 itself from O2 (Estevez et al., 1999) leading to spontaneous peroxynitrite

formation and protein nitrasylation (figure 2d). The peroxynitrite hypothesis was supported by

immnohistochemistry results from both mice (Andrus et al., 1998; Bruijn et al., 1997a;

Ferrante et al., 1997) and humans (Beal et al., 1997) showing elevated protein nitrotyrosine

levels as predicted by the hypothesis. However, in the same fashion as the loss of function

theory was disproved by showing that manipulation of SOD1 activity levels does not have a

positive effect on the disease progression applies also to the peroxynitrite theory, as increased

wild type SOD1 expression should quench the levels of superoxide and prevent them from

spontaneously forming peroxynitrite. As the toxicity of mutant SOD1 cannot be ameliorated

by increased SOD1 activity, the damage cannot be arising from superoxide or any

spontaneous reaction product of it.

Maybe the most attractive model for oxidative damage generated by aberrant substrate

is the peroxidase hypothesis. A second proposed aberrant substrate was hydrogen peroxide

(Wiedau-Pazos et al., 1996), which is interesting, as hydrogen peroxide is the normal end

product of the oxidized SOD1-Cu2+ form of the enzyme. Use of peroxide as a substrate by the

reduced SOD1-Cu1+ form might produce the extremely reactive hydroxyl radical (figure 2b).

An increase has been reported in the use of hydrogen peroxide by A4V- and G93A-SOD1

Page 35: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

35

Figure 2. SOD1 activity (a) and proposed aberrant substrates (b-d). a) Normal activity:SOD1 dismutases superoxide into oxygen and hydrogen peroxide in two asymmetric steps bythe reactive copper atom, which is alternately reduced and oxidized. Toxic hydrogen peroxideis converted to water by glutathione peroxidase or catalase. b) Hydrogen peroxide used as asubstrate by the reduced form of SOD1 may lead to formation of hydroxyl radicals. c)OONO as a substrate may lead to protein nitration. d) Zinc-depleted SOD1 may generate

superoxide from oxygen leading to perooxynitrite formation and further on to proteinnitration. Modified from (Cleveland and Rothstein, 2001).

SOD–Cu1+ SOD–Cu2+

O2•O2

H2O2H+ + •O2– 2H2O2

SOD–Cu1+ SOD–Cu1+–OH + OH–

H2O2

•OH

SOD–Cu1+ SOD–Cu2+

OH– + NO2-Tyr-Protein H-Tyr-Protein

•O2– + NO• –OONO

(Zn–) SOD–Cu1+ (Zn–) SOD–Cu2+ + •O2–

(Zn–) SOD–Cu2+

NO

NO2-Tyr-Protein (Zn–) SOD–Cu2+ + ONOO–

(reduced) (oxidized)

(reduced)

O2(oxidized)

a

b

c

d

Glutathione peroxidasecatalase

Page 36: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

36

mutants in vitro by spin trapping with 5, 5'-dimethyl-1-pyrrolline N-oxide (DMPO) (Wiedau-

Pazos et al., 1996).

However, the discovery was very soon challenged on technical grounds as it was

shown that a significant fraction of DMPO/•OH formed was derived from the incorporation of

oxygen from water due to oxidation of DMPO to DMPO/•OH, presumably via DMPO radical

cation (Singh et al., 1998). Still, products consistent with peroxidase hypothesis have been

found in the G93A-SOD1 mice strain as shown by increased lipid peroxidation (Hall et al.,

1998a) and increased protein carbonyl content (Andrus et al., 1998)

Overall, increased levels for markers of oxidative damage have been found in G93A-

SOD1 mouse strain (Andrus et al., 1998; Ferrante et al., 1997), but not in transgenic animals

with other mutations such as G37R-SOD1 (Bruijn et al., 1997a) or G85R-SOD1 (Williamson

et al., 2000). In human cases, markers of oxidative damage have been found in sporadic ALS

(Bowling et al., 1993; Ferrante et al., 1997; Shaw et al., 1995b) but not in SOD1-mediated

familial ALS (Bowling et al., 1993; Ferrante et al., 1997) and despite the elegant hypothesis

and hard work on the theories, aberrant substrate activity of SOD1 is most likely not the

primary cause of toxicity in ALS. Moreover, both of the theories and the proposed activities

require copper in the catalytic site of the enzyme and the strongest argument against the

theories comes from the discovery that an inactive SOD1 lacking all copper coordinating

histidines and also a reaction catalyzing copper ion still causes overt and progressive motor

neuron disease (Wang et al., 2003). On the other hand, this hypothesis does not account for

the possible formation of the hetrodimeric SOD1 (Witan et al., 2008) that might be formed in

the mouse models by mutant SOD1 and mouse endogenous wt SOD1, which is active and

contains copper binding sites. Furthermore, genetically decreased copper concentrations in

spinal cord of G86R-SOD1 mouse model of ALS have been shown to prolong life span by

9% (Kiaei et al., 2004), indicating a neurotoxic role for copper and active SOD1 in ALS

pathogenesis. Moreover SOD1 may affect ROS production, not just by its catabolic activity,

but by stimulating excess ROS production of NADDPH oxidase in microglia by binding to a

signaling GTPase protein Rac1 (Harraz et al., 2008).

2.2.2 Protein Aggregation

2.2.2.1 Aggregates

Protein aggregation is a common pathological hallmark of many different neurodegenerative

disorders like -amyloid and tau in Alzheimer's disease, -synuclein in Parkinsons's disease

Page 37: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

37

and huntingtin in Huntington's disease. For ALS, intracellular, cytosolic protein aggregates in

motor neurons and also in surrounding astrocytes are observed in both sporadic and familial

cases as well as in mutant SOD1 transgenic mice, where aggregates are highly

immunoreactive against SOD1 (Bruijn et al., 1997b; Bruijn et al., 1998; Watanabe et al.,

2001). In ALS the term aggregate has been used to refer to abnormal intermediate filaments

like neurofilaments and peripherin as detected by immunostaining of spinal cord tissue

(Hirano et al., 1984b), accumulation of detergent insoluble forms of proteins, including

SOD1, detected by immunoblotting (Johnston et al., 2000) as well as small SOD1 or ubiquitin

positive fibrillar inclusions in spinal cord sections (Wang et al., 2002). However, the thing is

that in ALS, as well as in other neurodegenerative diseases with protein aggregation, it is not

known whether the aggregates are toxic and by which mechanism, or are they a beneficial end

product of sequestered harmful and toxic intermediates of protein aggregates. Several

different theories have been put forward to explain the possible toxicity of protein aggragates,

including loss of protein function through coaggregation, depletion of protein folding

chaperones, dysfunction of the proteasome overhelmed by misfolded ubiquitinated protein

and inhibition of cell organelle function through aggregation within them e.g. in mitochondria

or peroxisomes (as reviewed in Bruijn et al., 2004b).

2.2.2.2 Proteasome and Immunoproteasome

Protein degradation is necessary for maintaining cellular homeostasis as cellular structures are

continually rebuilt and misfolded proteins, formed i.e. by mutations or oxidative stress must

be degraded. The ubiquitin proteasome pathway is the major proteolytic system of eukaryotic

cells, where it catalyzes the selective hydrolysis of ubiquitin tagged proteins in an ATP-

dependent manner (DeMartino and Slaughter, 1999).

Aggregates in patients and mouse models of ALS contain ubiquitin, a marker for the

protein in question to be degraded in the proteasome pathway. Excess accumulation of

ubiquitinated proteins may adversely affect the proteasome machinery and impair normal

proteasome function. For ALS, it has been shown that already partial inhibition of the

proteasome is sufficient to cause formation of large aggregates in cultured nonneuronal cells

that express mutant SOD1 (Johnston et al., 2000). However, from there on the results on

proteasome activity and function in ALS pathogenesis have been controversial as it has been

shown that overall proteasome activity is down-regulated in lumbar spinal cord well before

disease onset (Kabashi et al., 2004), upregulated at symptomatic stage (Puttaparthi and Elliott,

2005) or remains unaltered with disease progression (Cheroni et al., 2005). Though the results

Page 38: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

38

may seem contradictory, regulation of proteasome function may explain the differing results

on proteasome activity in ALS pathogenesis.

The proteasome is a 700 kDa cylinder shaped protease composed of four heptameric

rings of and subunits; the subunits form the outer rings and subunits form the inner

rings. In eukaryotes the subunits represent 14 gene products, seven of subunits and seven of

subunits. The proteasome is a multicatalytic protease charactherized by three specific

activities: a chrymotrypsin-like, trypsin-like and caspase-like activity and each catalytic

activity is linked with a specific subunit (as reviewed in Baumeister et al., 1998). In higher

eukaryotes each of the constitutive subunits, termed 1 or Y or ; 2 or Z; and 5 or X, has a

close homolog (LMP2, MECL-1 and LMP7, respectively) that can be selectively induced

with -interferon (Fruh et al., 1994), resulting in the replacement of their constitutive counter

parts and formation of proteasomes with altered catalytic characteristics that favor the

generation of peptides suitable for binding the MHC class I antigen presenting molecules

(Dick et al., 1996). In normal brains, the constitutive subunits are predominant, whereas the

inducible subunits are marginally expressed (Stohwasser et al., 2000).

Cheroni and colleagues showed for ALS for the first time that although overall

proteasome activity remained unaltered, the activity of constitutive proteasome decreases with

disease progression. This decrease was supplemented by induction of the immunoproteasome

so that the overall activity remained unaltered although levels of constitutive proteasome and

immunoproteasome varied (Cheroni et al., 2005). Similarly, a paper published later the same

year showed that immunoproteasome is activated in ALS, although this time changes with

constitutive proteasome levels were not observed. Hence the overall proteasome activity

increased with disease progression as immunoproteasome was induced (Puttaparthi and

Elliott, 2005). Even more importantly, Puttaparthi and Elliot showed that induction of

immunoproteasome was localized only to astrocytes and microglia, not to motor neurons.

This is interesting in the light of recent results showing that non-neuronal cells surrounding

motor neurons can either help the motor neurons to survive in the ALS pathogenesis, or to

strike a toxic cascade against them. Induction of the immunoproteasome would seem to be a

beneficial phenomenom that might help the motor neurons and non-neuronal cells to cope

with aggregating proteins.

2.2.2.3 Chaperones

Motor neurons have a high threshold for induction of the stress response which can contribute

to their selective vulnerability in ALS (Batulan et al., 2003). Keeping this is mind,

Page 39: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

39

destabilized and misfolded SOD1 is a possible initiator of aggregation through clogging of

protein folding chaperone machinery and heat shock proteins (HSPs). Multiple recombinant

SOD1 mutants inhibit chaperone function in vitro (Bruening et al., 1999) and even more

importantly, an overall decrease of chaperone activity has been reported in spinal cord

extracts of ALS mice from presymptomatic to the end stage of the disease (Tummala et al.,

2005). However, some chaperones are upregulated with disease progression; for example,

B-crystallin and Hsp27 are elevated in the spinal cords of G37R-SOD1 and G93A-SOD1

mice. Expression of B-crystallin has increased overall and in all cell types, but Hsp27 is

predominantly present only in glial cells and at late stages of the disease (Vleminckx et al.,

2002; Wang et al., 2003).

Elevating expression of different HSPs (Hsp70, Hsp40, Hsp27) in cultured cells and

primary motor neurons decreases aggregate content and apoptosis and improves axonal

outgrowth (Bruening et al., 1999; Patel et al., 2005; Takeuchi et al., 2002). However, this

positive effect does not apply in vivo, as elevated expression of Hsp70 to a level that is

protective in mouse models of acute ischemic insult and selective neurodegenerative disorders

did not ameliorate disease or pathology in four different mutant SOD1 lines (Liu et al., 2005).

2.2.2.7 Neurofilaments and axonal transport

Neurofilaments are the most abundant structural proteins in many types of mature motor

neurons. Aberrant neurofilament accumulations are a pathological hallmark of both familial

and sporadic ALS (Hirano, 1991; Hirano et al., 1984a; Hirano et al., 1984b) and are also seen

in mice expressing mutant SOD1 (Bruijn et al., 1997a; Dal Canto and Gurney, 1994).

Neurofilaments compose of heteropolymers of neurofilament-light (NF-L), -medium (NF-M)

and -heavy (NF-H) subunits. Transgenic mice expressing a mutation in NF-L develop motor

neuron disease as axonal architecture is perturbed (Lee et al., 1994), whereas deletion of NF-L

prolongs survival of two SOD1-mouse models of ALS (Nguyen et al., 2001; Williamson and

Cleveland, 1999), possibly because of the reduction of phospohorylated neurofilament tail

domains on the axons. Taken together, neurofilament content and disorganization are

important contributors and probable risk factors for the disease.

2.2.3 Glutamate excitotoxicity

Neurotransmitter glutamate released form presynaptic terminals triggers action potentials in

motor neurons by activating calcium-permeable AMPA glutamate receptors. If glutamate is

Page 40: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

40

not promptly cleared away from the synaptic cleft by glutamate transporters, chronic over

activation of glutamate receptors will cause repetitive neuron firing and increased intracellular

calcium. This glutamate mediated exitotoxicity can induce neuronal damage and death. Glial

glutamate transporter EAAT2 in astrocytes is one of the five subtypes of glutamate

transporters and is responsible for ~90% of the clearance of glutamate for motor neurons

(Rothstein et al., 1996; Tanaka et al., 1997). Preventing glutamate exitotoxicity within the

spinal cord is the job of astrocytes.

In ALS, glutamate excitoxicity involvement arose from the discovery of increased

glutamate levels in cerebrospinal fluid of ALS patients (Rothstein et al., 1991; Rothstein et

al., 1990; Shaw et al., 1995a) and is nowadays reported in 40% of sporadic ALS patients

(Spreux-Varoquaux et al., 2002). Measurement of functional glutamate transport in ALS

revealed an evident decrease in the affected brain regions caused by loss of glutamate

transporter protein EAAT-2 (Rothstein et al., 1995). Lowering of EAAT2 levels with an anti-

sense oligonucleotide has been shown to induce neuronal death both in vivo and in vitro

(Rothstein et al., 1996). In addition, expression of two SOD1 mutants in mice leads to

functional loss of EAAT2 (Nagano et al., 1996) and expression of G93A-SOD1 in rats

induces a focal loss of EAAT2 selectively from the astrocytes within the ventral horn of the

spinal cord (Howland et al., 2002). Motor neurons have a high sensitivity to excitotoxicity

since their AMPA glutamate receptors have a low portion of the calcium resistant GluR2

subunit, rendering them more permeable to calcium ions (Van Damme et al., 2002).

Glutamate excitotoxicity with astrocytic dysfunction is likely to be an important

contributor to neuronal death in ALS and is one of the few mechanistic links between

sporadic and mutant SOD1 mediated ALS. Moreover, the only currently FDA-approved

therapy in ALS, Riluzole, functions by decreasing glutamate toxicity (Bensimon et al., 1994;

Lacomblez et al., 1996).

2.2.4 Inflammation

Microglia are the resident immune cells of the central nervous system and are primary

mediators of neuroinflammation (Kreutzberg, 1996). In the resting state microglia have a

ramified shape characterized by a small cell body with fine processes and minimal expression

of surface antigens. They form a network of immune alert resident macrophages with a

capacity for immune surveillance and control. Upon injury to the CNS the microglia are

activated and acquire an ameboid shape. Activated microglia are mainly scavenger cells but

also perform various other functions in tissue repair and neural regeneration. Activated

Page 41: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

41

microglia can destroy invading micro-organisms, remove potentially deleterious debris,

promote tissue repair by secreting growth factors but they can also exert neuroinflammation

through the release of oxygen radicals, NO, cytokines, glutamate and prostaglandins. The

transformation of microglia into potentially cytotoxic cells is under strict control and occurs

mainly in response to neuronal or terminal degeneration, or both (Hanisch, 2002; Kreutzberg,

1996).

Microglial activation has been described in many acute and chronic neurological

diseases, including Alzheimer's disease (Kreutzberg, 1996; McGeer and McGeer, 1999). In

ALS patients, microglial activation and proliferation has been described in the brain and

spinal cord in regions of motor neuron loss (Engelhardt and Appel, 1990; Henkel et al., 2004;

Kawamata et al., 1992; McGeer et al., 1991; Troost et al., 1993; Turner et al., 2004).

Microglial activation is also seen in the spinal cord of different mutant SOD1 mice (Hall et

al., 1998b; Henkel et al., 2006; Kriz et al., 2002) where microglial reactivity is initiated before

motor neuron loss (Henkel et al., 2006) and expression of proinflammatory mediators like

TNF , IL-1 and COX-2 produced by microglia and other inflammatory cells is also an early

event (Alexianu et al., 2001; Almer et al., 2002; Elliott, 2001; Hensley et al., 2002; Nguyen et

al., 2001).

Microglial involvement in ALS is also supported by results showing that minocycline,

a tetracycline antibiotic derivative able to block microglial activation (Yrjänheikki et al.,

1999), also delays disease progression in ALS mice and reduces microglial activation (Kriz et

al., 2002; Van Den Bosch et al., 2002; Zhu et al., 2002). When minocycline treatment is

started in young presymptomatic mice, minocycline will delay disease onset but not disease

progression after the onset (Zhu et al., 2002), whereas administration at disease onset results

in slowing of the disease after onset (Kriz et al., 2002). However, later on minocycline failed

to show benefit in a large phase III humal trial of ALS (Gordon et al., 2007) and also findings

in the mouse model have been challenged (Scott et al., 2008). Inhibition of COX-2, expressed

by neurons and astrocytes, with celecoxib prolonged survival in mice by diminishing

asrtogliosis and microglial activation (Drachman et al., 2002), indicating the importance of

gliosis to the disease pathogenesis. However, also celecobix later failed to provide benefit in

human trials (Cudkowicz et al., 2006).

2.2.5 Mitochondria

Mitochondria among other cell organelles are likely targets for mutant SOD1 toxicity.

Although SOD1 is classically considered as a cytosolic enzyme (Fridovich, 1986), it is also

Page 42: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

42

located in the mitochondrial intermembrane space (Okado-Matsumoto and Fridovich, 2001;

Sturtz et al., 2001). Vacuolated and dilated mitochondria with disorganized cristae and

membranes have been observed in ALS early on in muscles and spinal cord of both sporadic

and familial patients (Afifi et al., 1966; Hirano et al., 1984a; Hirano et al., 1984b). Since then

it has been shown with mutant SOD1 animal models that mutant SOD1 is preferentially

imported into mitochondria in affected tissues (Bergemalm et al., 2006; Deng et al., 2006; Liu

et al., 2004; Vijayvergiya et al., 2005). If we consider the hypothesized theories of mutant

SOD1 toxicity on mechanisms leading to motor neuron cell death, including oxidative

damage, aggregation and excitotoxicity, it can be noted that mitochondria can be directly

linked to all of these harmful cascades in ALS pathogenesis. However, there is yet no specific

mechanism for the toxicity of SOD1 in mitochondria and many contradictions remain to be

resolved.

Oxidative stress is a likely cause of mitochondrial damage as mitochondria themselves

are a source of reactive oxygen species as an unavoidable byproduct of oxidative respiration.

Mutant SOD1 may interfere with the electron transport chain disrupting ATP production and,

in fact, ATP levels have been shown to be depleted in presymptomatic G93A-SOD1 mice

(Browne et al., 2006) and at the time of onset, ATP synthesis itself was defective (Mattiazzi et

al., 2002). However, this holds only for the G93A-SOD1 model as ATP synthesis has been

shown to be unaffected in aged symptomatic G85R-SOD1 mice (Damiano et al., 2006), and

also other aspects of electron transport chain have been reported as either unchanged or

altered. Activity of the mitochondrial cytochrome oxidase has been shown to be elevated

(Bowling et al., 1993), reduced (Mattiazzi et al., 2002) and unchanged (Damiano et al., 2006).

In addition, treatment with creatine extended the survival in G93A-SOD1 mice alleviating

energy deficits (Browne et al., 2006; Klivenyi et al., 1999), but did not have any beneficial

effects in human clinical trials (Groeneveld et al., 2003; Shefner et al., 2004).

One of the most convincing connections between SOD1 toxicity and mitochondria is

the disruption of mitochondrial calcium buffering capacity in SOD1 mice. Although studies

on mitochondrial energy metabolism had no consensus of the mutant SOD1 action between

different mouse models or human ALS, calcium-buffering capacity has been shown to be

impaired in spinal cord of both G93A-SOD1 and G85R mice at presymptomatic stage and

only in affected tissues (Damiano et al., 2006). Mutant SOD1-mediated disruption of calcium

homeostasis is also compatible and can be linked directly with excitotoxic mechanisms

leading to neuronal cell death as, excitotoxicity from repeated neuron firing increases

intracellular calcium levels and leads to the opening of mitochondrial membrane permeability

transition pore (PTP) and release of cytochorme c triggering the apoptotic process. Caspase

Page 43: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

43

mediated apoptotic neuronal death and activation of caspase 3 is the final event in ALS

pathogenesis appearing in motor neurons and astrocytes (Li et al., 2000; Pasinelli et al.,

2000). Moreover, activation of caspases can be linked back to excitotoxicity as caspase 3

activation in glial cells proteolytically inactivates the glutamate transporter enhancing

excitotoxicity, calcium entry and mitochondrial damage (Boston-Howes et al., 2006).

Mutant SOD1 aggregates may also damage mitochondria and induce apoptosis by

interfering with anti-apoptotic activity of Bcl-2 and triggering the premature activation of the

apoptotic process (Pasinelli et al., 2004). Mutant SOD1 aggregating onto the mitochondrial

surface might also impede the transport of nuclear encoded proteins into mitochondria

through translocators of the outer and inner membrane (TOM and TIM, respectively) (Liu et

al., 2004). In addition to aggregates, high levels of mutant SOD1 expression and

mitochondrial abnormalities have been detected in motor neuron presynaptic terminals at the

onset of neuromuscular denervation, before accumulation of mutant SOD1 (Gould et al.,

2006).

Although SOD1 is believed to be an important part of the mitochondrial superoxide

scavenging system (Sturtz et al., 2001), the role of SOD1 in the mitochondria is somewhat

controversial. In the mitochondrial matrix superoxide is dismutated to oxygen and hydrogen

peroxide by Mn-superoxide dismutase (SOD2) (Zelko et al., 2002) and in the intermembrane

space superoxide detoxification is thought to depend on cytochrome c, which can efficiently

oxidize superoxide to oxygen acting as a true antioxidant (Pereverzev et al., 2003), whereas

SOD produces oxygen and hydrogen peroxide that further needs to be detoxified by

glutathione peroxidase or catalase, which are present in the intermembrane space only in low

levels (Martin et al., 1998). Moreover, SOD1 in the intermembrane space is mainly kept

inactive (Inarrea et al., 2005) raising questions about the suggested role in the mitochondrial

superoxide scavenging system.

2.2.6 Role of non-neuronal cells

In the previous chapters the role of microglia as the resident immune cell of the central

nervous system and the important function of astrocytes in clearing the neurotransmitter

glutamate from the synaptic cleft of motor neurons were described briefly. At this point it is

important to note that although ALS affects only motor neurons, the disease pathogenesis is

not, however, cell-autonomous. In other words, the toxicity that kills the motor neurons does

not rise solely from the mutant SOD1 expressing motor neurons themselves, but also from the

Page 44: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

44

surrounding non-neuronal cells as astrocytes also show pathology (Bruijn et al., 1997b) and

microglia are activated in the regions of motor neuron loss (Hall et al., 1998b).

Restricted expression of mutant SOD1 in astrocytes (Gong et al., 2000) or motor

neurons alone (Lino et al., 2002; Pramatarova et al., 2001) failed to provoke overt motor

neuron disease in mice, supporting the idea of non-cell- autonomous disease pathogenesis, as

reviewed in chapter 2.1.4.1. Moreover, chimeric mice having both mutant SOD1 and normal

cells in their CNS showed that even a fraction of normal cells in the CNS can save the

animals from the disease, despite high levels of mutant SOD1 in motor neurons (Clement et

al., 2003).

The role of mutant SOD1 toxicity in microglia and in motor neurons themselves was

further analyzed by the Cre/Lox approach, where mutant SOD1 gene flanked by LoxP sites

was deleted from microglia or motor neurons by Cre-recombinase excision targeted either to

motor neurons by a motor neuron specific promoter Islet-1, or to microglia and peripheral

macrophages with a CD-11b promoter (Boillee et al., 2006b). This approach has identified the

differential contribution of mutant damage within motor neurons and microglia on disease

onset and progression. Excision of mutant SOD1 within motor neurons extended survival of

the mice by slowing onset and early phase of disease progression. Similar effects on survival

have been obtained by removal of neurofilaments (Williamson et al., 1998) or the tail domain

of the NF-M and NF-H subunits (Lobsiger et al., 2005) specifically in motor neurons. The

deletions provided extended survival but only by slowing disease onset. In contrast,

diminishing mutant SOD1 expression in microglia has little effect on the onset of disease but

slows later disease progression, resulting in remarkable extension of the overall survival as

the increase in survival was longer for Cre-CD11b animals than for Islet-1 mice. In addition

to floxing of mutant SOD1 form microglia and slowing down late phase of the disease,

similar results have been obtained in a complementary study, where the entire myeloid

lineage was replaced by transplantation of normal bone marrow cells into SOD1 mutant mice

that had a deletion of PU.1 transcription factor, making them unable to produce myeloid cells.

Transplantation had no effect on onset but slowed disease progression after the onset (Beers et

al., 2006). Moreover, transplantation of mutant G93A-SOD1 myeloid cells at birth into

G93A-SOD1 / PU.1 knock out mice produced onset and survival typical of the G93A-SOD1

line, whereas transplantation of G93A-SOD1 bone marrow cells into wild type animals did

not give rise to motor neuron disease, demonstrationg that mutant SOD1 in microglia alone is

not sufficient to cause motor neuron death, but accelerates disease progression after the onset.

Mutant SOD1 in astrocytes can also affect the survival of motor neurons in vitro.

Cultured mouse motor neurons from embryonic spinal cord or differentiated from embryonic

Page 45: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

45

stem cells were less likely to survive when co-cultured with mutant SOD1 expressing

astrocytes (Di Giorgio et al., 2007; Nagai et al., 2007). Interestingly, in culture conditions the

survival of motor neurons was not affected by coculture with mutant SOD1 expressing

microglia or fibroblats. Moreover, astrocytes expressing mutant SOD1 did not show toxic

effects on other neuronal cells, such as GABAergic neurons or dorsal root ganglion neurons,

indicating a selective but unknown vulnerability of motor neurons (Nagai et al., 2007). But,

what is the toxic mechanism of astrocytes? Excitotoxicity through decreased glutamate uptake

by astrocytes is a central toxic mechanism implicated in ALS and glutamate transporter levels

are reduced in mutant SOD1 animals. However, extracellular glutamate was not increased in

the culture system (Nagai et al., 2007). Activated astrocytes also produce cytokines,

chemokines and other neurotoxic substances that may mediate neuronal death. However, no

abnormalities were observed in IL-1 , IL-6, INF- , TNF- or FAS ligand (Nagai et al., 2007).

The toxic mechanism remains to be elucidated, yet reactive oxygen species produced by

astrocytes might mediate motor neuron death. Secreted mutant SOD1 might be also the

damaging component directly as mutant SOD1 may be secreted by astrocytes. Extracellular

mutant SOD1 was recently shown to provoke motor neuron cell death in cultures and it is a

potent activator of microglial cells (Urushitani et al., 2006).

Summarizing the role of non-neuronal cells and motor neurons; it seems that mutant

SOD1 in glia accelerates disease progression considerably although it cannot induce disease

alone, while mutant action within motor neurons is a primary determinant of early onset

disease. It also extremely interesting to note that when ALS mice are treated with minocycline

starting from a very young age, minocycline delays disease onset (Zhu et al., 2002), while

administration of minocycline closer to the disease onset results in a slowing of disease

progression after onset (Kriz et al., 2002), suggesting that minocycline may have an effect on

both glia and neurons alleviating the toxicity. The role of non-neuronal cells has also

implications on stem cell therapies for ALS, as it may not be a good idea to transplant motor

neurons to a hostile glial environment, but rather transplant healthy astrocytes or microglia.

2.2.7 Pathway of motor neuron cell death in ALS

ALS is a multifactorial complex disease with many proposed mechanism leading to specific

cell death of motor neurons. However, instead of looking to proposed mechanisms i.e.

aggregation, excitotoxicity, oxidative damage or neurofilament disorganization as separate

causes or trying to determine which is the primary one, the hypothesized mechanisms can be

viewed as a combined pathogenic pathway (Figure 3) (as reviewed in Boillee et al., 2006a).

Page 46: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

46

The following is a description of the pathogenic pathway from the onset to the end stage in

the inherited form of ALS composed from the results obtained form mutant SOD1 mouse

lines. Although only motor neurons are degenerating, it is important to note that damage is

not rising from the motor neurons alone but also from the surrounding non-neuronal cells. On

the other hand, one may ask why only motor neurons are degenerating despite ubiquitous

expression of mutant SOD1. The answer is because of unique functional properties of the

motor neurons; they are very large cells having axons over one meter in length in humans

reaching form spinal cord to feet. This sets high demands for energy production and

biosynthesis. Motor neurons have high rates of firing and response to glutamate inputs,

making them efficient but vulnerable to excessive glutamate and calcium.

As shown by Boillee et al. 2006b, damage within motor neurons is an important

determinant of disease onset. The earliest event in human ALS and in mutant SOD1 mice is

denervation, or in other words, withdrawal of motor neurons axons from their synapses on

muscles (Fischer et al., 2004; Frey et al., 2000; Pun et al., 2006; Schaefer et al., 2005). This

denervation causes the first symptom in ALS pathogenesis that is visible to the eye as seen by

loss of strength from the voluntary muscles that will progress to paralysis and muscle atrophy

(Cifuentes-Diaz et al., 2001). During the phase of denervation at the time of onset of the

disease, mutant SOD1 primarily acts within motor neurons as misfolded and aggregating

SOD1 damages cellular machinery, especially mitochondria, proteasome and chaperones,

inhibiting their normal functions in maintaining cellular homeostasis. Axonal transport is also

inhibited as a consequence of neurofilament disorganization, blocking the transit of cellular

components, i.e. mitochondria, to sites where energy is needed or resulting from

mitochondrial damage as energy needed for the transport is no longer supplied to the transport

machinery. The action of mutant SOD1 in motor neurons is also amplified by action within

other cell types, especially in microglia, as they respond to the initial damage through positive

feedback by activation and proliferation.

The symptomatic phase is characterized by massive gliosis or activation of microglia

and astrocytes in addition to continuing damage within motor neurons. Mutant SOD1 can

itself activate microglia, promoting secretion of trophic molecules but also the secretion of

toxic proinflammatory factors like nitric oxide, TNF and IL-1 produced by activated

microglia and astrocytes, which activates neighboring cells and thus end up harming their

environment and motor neurons. Toxic mutant SOD1 may also be secreted or released

through cell leakage or lysis. Mutant SOD1 in astrocytes induces loss of the EAAT2

glutamate transporter, causing repetitive firing of glutamate receptors and excessive influx of

calcium on the motor neurons as glutamate is not rapidly cleared away from the synaptic cleft.

Page 47: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

47

Figure 3. Mechanisms leading to motor neuron degeneration in ALS. Toxicity arises fromdamage within motor neurons and is amplified by activated microglia and astrocytes.Selective vulnerability of motor neurons to mutant SOD1 toxicity is resulting from theirunique properties (large cells with long axons, high biosynthetic load, high rates of firing andglutamate uptake) and damage to supporting non-neuronal cells. Modified from (Boillee et al.,2006a; Cleveland and Rothstein, 2001)

Ca2+

ActivatedMicroglia

ActivatedAstrocyte

Motor Neuron

Muscle

Schwann Cell

GlutamatergicNeuron

SOD1

Mitochondria

Proteasome

Chaperones

Endosomal reticulum

Caspase

Neurofilament Disorganization

EAAT2 is lostToxic factors(e.g. NO, NGF)

Glu

Toxic factors(e.g. NO, TNF )

EAAT2

Denervation

Aggregates containing SOD1

Page 48: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

48

At the end stage, denervation has produced paralysis of voluntary muscles

accompanied by muscle atrophy. Activated microglia and astrocytes continue to produce

diffusible toxic products including nitric oxide and TNF , accelerating disease spread.

Repetive firing of glutamate receptors from lack of EAAT2 result in excitotoxicity with

excessive intracellular calcium that further damages calcium buffering mitochondria and

endosomal reticulum and triggers the caspase dependent programmed cell death pathway

within motor neurons. To summarize, the selective degeneration of motor neurons is not alone

due to their unique and possible fragile form and functions, but from a combination of factors

rising from motor neurons and the surrounding glia.

2.3 Therapeutics for ALS

2.3.1 Drug treatments

ALS is a complex multifactorial disease, making the discovery of effective drugs and

interventions challenging. On the other hand, as many different pathogenic mechanisms are

involved in ALS, intervention on any of these mechanisms may interrupt or slow down the

pathogenesis. However, no cure or effective treatment presently exists for ALS.

Many different types of drugs have been tested in rodent models of ALS and also in

clinical trials and most are based on various hypotheses of mechanisms for neuronal death

that are learned by studying the mutant SOD1 mice. The treatments have been targeted to

overcome e.g. oxidative damage, loss of trophic factor support, glutamate-mediated

excitotoxicity and chronic inflammation (Bruijn et al., 2004a). Riluzole, an inhibitor of

glutamate release and an anti-convulsant, is so far the single agent presently approved for

clinical use. It only extends survival modestly by a few months but has little or no effect on

the symptoms (Bensimon et al., 1994; Lacomblez et al., 1996). Paradoxically, it may itself

cause muscle weakness although these adverse effects are reversible after stopping the use of

the drug. A number of trophic factors, anti-inflammatory agents, and inhibitors of oxidative

stress have been reported to prolong survival in mouse models and some are now in clinical

trials (Bruijn et al., 2004a). VEGF, anti-inflammatory COX-2 inhibitor celecoxib, and

minocycline have had particularly promising results in mice (Azzouz et al., 2004; Cudkowicz

et al., 2006; Kriz et al., 2002; Storkebaum et al., 2005; Van Den Bosch et al., 2002; Zhu et al.,

2002). Immunization against mutant SOD1 may also have therapeutic effects as mutant SOD1

has secretory pathways and secreted extracellular mutant SOD1 can provoke motor neuron

cell death in culture (Urushitani et al., 2006). Passive immunization through intraventicular

Page 49: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

49

infusion of anti-human SOD1 antibody alleviated symptoms and increased life span of G93A-

SOD1 mice (Urushitani et al., 2007), making immunization one possible candidate for the

treatment of familial ALS caused by SOD1 mutations.

Clinical trials for several of these treatment strategies are ongoing, but no

breakthroughs have yet occurred, whereas a rather dissapointing moment was faced recently

when minocycline failed to show any benefit in a large phase III humal trial of ALS (Gordon

et al., 2007). Presently it is thought that combinations of different drugs may be required to

slow the multifactorial neurodegeneration process effectively (McGeer and McGeer, 2005).

2.3.2 Growth factors

Growth factor therapies on ALS have been based on the hypothesis that whatever the disease

provoking insult is, providing increased levels of trophic factors to motor neurons would be

beneficial. However, so far clinical trials with growth factors have had only little or no

benefit. Insulin like growth factor 1 (IGF-1) slowed the progression of functional impairment

and the decline in health-related quality of life in patients with ALS in one study (Lai et al.,

1997), but failed to do so in another (Borasio et al., 1998). No beneficial effects were seen

with trials on BNDF or CNTF either (ALS CNTF Treatment Study Group 1996; BDNF Study

Group 1999). One major problem in these trials has definitely been the delivery of factors past

the blood-brain-barrier to the CNS, as growth factors were delivered subcutaneously or

directly into the spinal fluid. Direct intracerebroventicular (ICV) or intrathecal infusion of

growth factors might overcome the delivery problem. Although intrathecal administration of

BDNF was found to have no effect, ICV infusion of IGF-1 extended survival in mice (Nagano

et al., 2005a) and also showed a modest benefit to ALS patients in a clinical trial (Nagano et

al., 2005b).

In addition to direct infusion, growth factors have been delivered in mouse models

with viral vector mediated gene therapy. Adeno-associated virus (AAV) has been used to

retrogradedly transport the growth factor gene along the axon from muscle to the motor

neuron cell soma where the viral genome can express the growth factor gene and trophic

factors are secreted by the motor neuron. Delivery of IGF-1 with this strategy slowed disease

progression in mice even when the treatment was initiated after disease onset (Kaspar et al.,

2003).

VEGF gene may be a contributor to ALS, as mutations of VEGF gene and its

promoter are associated with increased risk of developing ALS (Lambrechts et al., 2003;

Terry et al., 2004). Therefore, both viral and ICV delivery of VEGF have been tested in ALS

Page 50: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

50

models. Muscle injected with lentiviral VEGF was retrogradedly transported to motor neurons

and extended survival of ALS mice (Azzouz et al., 2004), as did also continous ICV infusion

of recombinant protein into CSF in ALS rats (Storkebaum et al., 2005).

2.3.3 Gene therapies

As mutant SOD1 causes ALS through a toxic gain of function and lack of SOD1 activity in

knock out mice is not detrimental, limiting mutant SOD1 expression either through viral

delivery of silencing RNA (siRNA) or direct infusion of antisense oligonucleotides might

offer useful treatment strategies.

In mice, lentiviral delivery of SOD1 siRNA through retrograde transport to motor

neurons reduced SOD1 expression in motor neurons, slowed disease initation and increased

survival remarkably (Ralph et al., 2005). However, disease progression from onset to end

stage was not affected at all as SOD1 was not silenced in surrounding non-neuronal cells.

A similar effect was also seen in another study using AAV as a vector (Miller et al., 2005).

Intrathecal injection of SOD1 siRNA virus to spinal cord also delayed motor neuron

degeneration near the injection site, but it had no effect on the overall disease progression

(Ralph et al., 2005).

However, when considering human clinical trials, some practical issues remain. For

example the dosage cannot be altered nor can the treatment be stopped as the virus starts the

expression of the delivered gene. An alternative to the problems of using viral vectors might

be delivery of antisense oligonucleotides. This approach has been succesful in ALS rats

resulting in lowering of SOD1 levels in brain and spinal cord and slowing of disease

progression (Smith et al., 2006).

2.3.4 Stem cell therapies

Selective cell death of motor neurons has made the idea of replacing dying motor neurons

with stem cells tempting. However, in ALS it is hard to imagine that even if transplanted stem

cells developed into motor neurons, how they can form appropriate connections with muscles

as motor axons need to extend distances for up to a meter in length to reach the target in

humans. Nevertheless, stem cell therapy for replacing motor neurons has been tried with some

success in Sindbis virus paralyzed rat spinal cords, where ES-cell -derived motor neuron

precursor cells were injected into rat spinal cord (Deshpande et al., 2006; Harper et al., 2004).

In the study by Harper et al. the cells survived and even extended axons with the help of

Page 51: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

51

intrathecal injection of myelin repulsion alleviating molecules, but no neuromuscular

junctions were formed (Harper et al., 2004). Altered differentiation conditions, dibutyryl as an

anti-myelin repulsion factor and grafting of neural stem cells expressing glial derived

neurotrophic factor (GDNF) produced neuromuscular connections and relieved paralysis of

Sindbis virus exposed rats (Deshpande et al., 2006).

An alternative strategy to the task of replacing the long motor neurons would be to

replace surrounding non-neuronal cells, as mutant SOD1 expressing motor neurons can be

saved by surrounding wild type glial cells and extend survival of motor neurons that express

SOD1 (Boillee et al., 2006b; Clement et al., 2003). Then also other sources of stem cells

instead of motor neuron precursor cells such as bone marrow or umbilical cord blood could be

used. Human umbilical cord blood cells have also been shown to be protective (Garbuzova-

Davis et al., 2003). Bone marrow transplantations have been used with G93A-SOD1 mice in

two instances, where the first reported paper showed some extension in survival (Corti et al.,

2004) whereas in the other no protection was seen (Solomon et al., 2006). The hurdle may be

the surviving residential microglia expressing mutant SOD1. Transplantation of normal bone

marrow cells into SOD1 mutant mice that had a deletion of PU.1 transcription factor (making

them unable to produce myeloid cells) slowed disease progression after the onset (Beers et al.,

2006). Moreover, transplantation of mutant G93A-SOD1 myeloid cells at birth into G93A-

SOD1 / PU.1 knock out mice produced onset and survival typical of the G93A-SOD1 line

whereas transplantation of G93A-SOD1 bone marrow cells into wild type animals did not

give rise to motor neuron disease. These results demonstrate that mutant SOD1 in microglia

alone is not sufficient to cause motor neuron death, but expression of mutant SOD1 in

microglia rather accelerates disease progression after the onset (Beers et al., 2006).

Accordingly, bone marrow replacement might be a beneficial treatment for ALS if the

recruitment to the brain is sufficient enough to replace the residential mutant SOD1

expressing microglia possibly even if the treatment is done after the onset.

In addition to replacing cells, stimulation of the endogenous pool of stem cells to

proliferate and replace degenerating motor neurons might provide neuroprotection. In G93A-

SOD1 mice some neurogenesis has been observed in response to neurodegeneration (Chi et

al., 2006). However, the proliferating endogenous cells also express mutant SOD1 and

therefore are susceptible to the mutant SOD1 toxicity.

Page 52: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

52

2.3.5 Pyrrolidine dithiocarbamate (PDTC)

Pyrrolidine dithiocarbamate (PDTC) belongs to a class of dithiocarbamates which have

previously been used in the treatment of bacterial and fungal infections, and have been

considered for use in the treatment of AIDS (Reisinger et al., 1990). PDTC is also known as

an inhibitor of nuclear transcription factor kappa-B (NF- B) that regulates the expression of

several proinflammatory genes and some genes related to apoptosis (Hayakawa et al., 2003;

Liu et al., 1999; Schreck et al., 1992). In addition, PDTC has been shown to be a potent

antioxidant both in vitro and in vivo (Hayakawa et al., 2003; Liu et al., 1999; Schreck et al.,

1992).

As inflammation, apoptosis and oxidative stress are implicated in a large range of

diseases including ALS, it is not surprising that PDTC has been reported to have beneficial

effects in models of diseases such as pleurisy, arthritis, colitis, liver and brain ischemia, spinal

cord injury, Alzheimer's disease, Duchenne muscular dystrophy, abdominal aortic aneurysm,

neonatal asphyxia and autoimmune uveoretinitis (Chen et al., 2005; Cheng et al., 2006;

Cuzzocrea et al., 2002; Kitamei et al., 2006; La Rosa et al., 2004; Malm et al., 2007; Matsui et

al., 2005; Messina et al., 2006; Nurmi et al., 2006; Nurmi et al., 2004a; Parodi et al., 2005).

There is evidence that mechanisms of PDTC's beneficial effects in these models are indeed

related to it's ability to act as an antioxidant and to inhibit the expression of pro-inflammatory

genes, including COX-2, TNF and interleukin-1 (Nurmi et al., 2004a; Nurmi et al., 2004b).

In addition, PDTC may also provide protection by acting through Akt kinase -

glycogen synthase kinase (GSK) signaling (Malm et al., 2007; Nurmi et al., 2006). Activation

of GSK-3 can lead to apoptotic neuronal death and result in energy depletion in stress

conditions. In addition, GSK-3 may also inhibit the expression of transcription factors that

support cell survival (Grimes and Jope, 2001). GSK-3 is a unique enzyme in the regard that

it is activated by dephosphorylation instead of the more common way of enzyme activation

trough phosphorylation. Akt is a well known kinase able to phosphorylate GSK-3 and thus

inactivate GSK-3 (Grimes and Jope, 2001). As PDTC activates Akt, it not surprising that

PDTC was shown to be protective in models of neonatal asphyxia and Alzheimer’s disease

with GSK-3 activation (Malm et al., 2007; Nurmi et al., 2006). PDTC can act also as a metal

chelator for transitional metals transporting extracellular copper into the cell and, moreover,

the transitional metal-PDTC complex may have an inhibitory activity against the proteasome

(Kim et al., 2004).

Page 53: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

53

PDTC has been shown to have beneficial effects in many different disease models and

it possesses capabilities for activating or inhibiting several cellular targets, making it an

interesting drug candidate for preclinical testing in ALS.

Page 54: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

54

3. AIMS OF THE STUDY

The underlying initiative cause of the mutant SOD1 toxicity is still unknown. Mutations of

SOD1 may affect the stability of the enzyme making it more prone to aggregate, mutations

may alter the catalytical site allowing aberrant substrates to enter the catalytical site and cause

unwanted oxidative reactions. Recent results have shown that the toxic effect of mutant SOD1

may be targeted to mitochondria, but the evidence for the toxic mechanism in mitochondria is

scarce. This study was carried out to assess the stability and oxidation of human mutant SOD1

throughout the disease progression using transgenic ALS rats and mice and to analyze the role

of mutant SOD1 in mitochondria. In addition, as oxidative damage, inflammation and

apoptosis play major roles in the pathology of ALS, anti-inflammatory and anti-oxidative

drug treatment with PDTC might have neuroprotective effects, and we tested PDTC drug

treatment on G93A-SOD1 transgenic rats. The specific aims were to study:

1) Effect of an anti-inflammatory and anti-oxidative drug treatment with PDTC on G93A-

SOD1 transgenic rats and analysis of the molecular mechansims involved.

2) Stability and molecular features of mutant SOD1 in a G93A-SOD1 rat model of ALS and

the effect of mutant SOD1 from affected tissues on the mitochondria.

3) Molecular mechanism for the deleterious role of mutant SOD1 in mitochondria.

4) Proteomic analysis of differential protein expression and protein oxidation in a G93A-

SOD1 mouse model of ALS with emphasis on oxidation of mutant SOD1 through disease

progression in affected and unaffected regions of the central nervous system.

Page 55: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

55

4. MATERIALS AND METHODS

4.1 Animals (I-IV)

All animal studies were carried out with permission of The Institutional Animal Care and Use

Committee of the University of Kuopio and the Provincial Government according to the

National Institute of Health guidelines for the care and use of laboratory animals. Rats and

mice were housed in groups of 2-3 rats in one cage in light and temperature controlled

environment with ad libitum water and standard laboratory rodent chow.

Transgenic rats expressing human mutant G93A-SOD1 [Tac:N:(SD)-

TgN(SOD1G93A)L26H, Emerging Models, sponsored by Amyotrophic Lateral Sclerosis

Association, Taconic, Hudson, NY, USA] develop motor neuron disease similar to human

ALS (Howland et al., 2002). Hemizygous rats over-express human mutant G93A-SOD1 at

levels increasing from 8-fold over endogenous SOD1 in presymptomatic rats to 16-fold in end

stage animals. Disease onset is determined by limb gait and occurs on average at 115 days (16

weeks). From the onset the disease progresses rapidly to end stage within 11 days. The end

stage of the disease was determined by righting reflex test, where the animal was placed on its

side and if the rat was not able to right itself in 30 seconds, it was scored as death and

sacrificed. Righting reflex failure typically coincides with complete paralysis of both

hindlimbs and one forelimb as a result of substantial loss of spinal cord motor neurons, as

well as marked increases in gliosis and degeneration of muscle integrity and function.

Transgenic and corresponding wild-type littermate rats were sacrificed at 8 weeks

(presymptomatic), 16 weeks (onset) and at the end stage of the disease (18 weeks of age).

Sacrificed animals were anaesthetized with 40mg/kg pentobarbital and perfused transcardially

with saline and extracted tissues were frozen in liquid nitrogen and stored at –80 C. G93A-

SOD1 rats were also treated with PDTC on a dose of 50 mg/kg/day starting at 70 days of age

(n=19). The control tg rats (n=12) received plain water. In addition, wt littermates of

corresponding age received either PDTC (n=6) or plain water (n=7). PDTC treated or

untreated animals were anaesthesized with 40mg/kg pentobarbital, perfused transcardially

with saline and extracted tissues were either frozen in liquid nitrogen and stored at –80 C or

post-fixed with 4% paraformaldehyde for 24 h, cryoprotected with 30% sucrose for three days

and then frozen in liquid nitrogen and stored at –80 C.

Transgenic G93A-SOD1 mice [B6.Cg-Tg(SOD1-G93A)1Gur/J, The Jackson

Laboratory, Bar Harbor, ME, USA] express high levels of the transgene with a 4-fold increase

in SOD activity, and exhibit a phenotype similar to ALS in humans (Gurney et al., 1994).

Page 56: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

56

Hemizygous transgenic mice become paralyzed in one or more limbs and have a life span of

approximately 19-23 weeks. Paralysis is due to loss of motor neurons from the spinal cord.

Mice were sacrificed with carbon dioxide anaesthesia and decapitation at the age of eight

weeks and tissues (spinal cord, brain and liver) were collected for mitochondria isolation.

Samples from G93A-SOD1 mice at ages of 8-, 14- and 18-weeks and corresponding wt

control mice and mice expressing human wt SOD1 age of 12 weeks were kindly provided by

Dr. Catherine Bendotti from the Department of Neuroscience, Istituto di Ricerche

Farmacologiche Mario Negri, Milan, Italy.

SOD1-/- knockout mice (Reaume et al., 1996) and corresponding wild types (n = 5)

were kindly provided by Dr. Pak Chan from the Department of Neurosurgery, Stanford

University School of Medicine, Stanford, California, USA. Knockout mice were sacrificed

with carbon dioxide anaesthesia and decapitation at the age of eight weeks and tissues (spinal

cord, brain and liver) were collected for mitochondria isolation. Prior to decapitation blood

was collected with heart puncture for lymphocyte isolation.

Wistar rats (8 week old males, National Laboratory Animal Center, University of

Kuopio, Kuopio, Finland) were used as source of liver mitochondria. Wistar rats were

sacrificed with carbon dioxide anaesthesia and decapitation.

4.2 PDTC treatment (I)

G93A-SOD1 rats were treated with PDTC on a dose of 50 mg/kg/day starting at 70 days of

age (n=19). Fresh PDTC (Sigma, St. Louis, MO, USA) was administered into drinking water

every other day. The control tg rats (n=12) received plain water. In addition, wt littermates of

corresponding age received either PDTC (n=6) or plain water (n=7). Neurological signs of

disease, such as overall locomotor activity, ataxia, extension reflex of hind limbs, limb gait,

paralysis of hind limb or paralysis of fore limb and weight gain of the animals were followed

daily. Appearance of limb gait was determined as the onset of the disease. A new dose of

PDTC was calculated weekly according to animal weight and water consumption, and the

treatment was continued until the end stage of the disease when the animals were sacrificed.

The end stage of the disease was determined by rightening reflex test where animal was

placed on its side and if it was not able to righten itself in 30 seconds it was scored as death.

Sacrificed animals were anaesthesized with 40 mg/kg pentobarbital, perfused transcardially

with saline and extracted tissues were either frozen in liquid nitrogen and stored at –80 C, or

post-fixed with 4% paraformaldehyde for 24 h and cryoprotected with 30% sucrose for three

days and then frozen in liquid nitrogen and stored at –80 C.

Page 57: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

57

4.3 Mitochondria (II,III)

4.3.1 Isolation of mitochondria

Mitochondria were isolated from liver, brain or spinal cord. Tissue was homogenized with all-

glass dounce homogenizator (Kontes, Vineland, NJ, USA) in 10 × volume of ice-cold

isolation buffer (320 mM sucrose, 1 mM EGTA, 10 mM Tris-HCl pH 7.4). The homogenate

was centrifuged at 2000 g for 3 min at 4 C, and the supernatant was transferred to a new

tube and centrifuged 10000 × g for 10 min. The supernatant was discarded and the pellet

containing the crude mitochondrial fraction was washed once with wash buffer (0.2 M

sucrose, 20 mM HEPES pH 7.2, 0.1 mM EGTA, 4 mM KH2PO4) and resuspended in the

same buffer to a concentration of 10 mg/ml. The protein concentration was determined by

protein assay dye (Bio-Rad, Hercules, CA, USA).

In order to obtain pure mitochondria from CNS tissue, the crude mitochondrial

fraction was transferred on top of a 19 % percoll solution (Sigma) in isolation buffer and

centrifuged at 30700 g and at 4 C for 10 min. The mitochondrial fraction accumulating to

the bottom was collected, diluted 1:3 with wash buffer, centrifuged at 17000 × g for 10 min at

4 C and resuspended back to wash buffer. The purity of the mitochondrial fraction was

determined with Western blotting as a presence of mitochondrial marker COX4 and absence

of cytosolic marker actin (figure 4).

Figure 4. Purity of mitochondrial fraction was determined with Western blotting as apresence of mitochondrial markers COX4 and SOD2 and absence of cytosolic marker actin.Subcellular fractions and markers from mouse cortex: Cyt = Cytosolic soluble fraction; Tot =Total homogenate; P2 = Crude mitochondrial fraction; Mito = Purified mitochondrial fraction.

MWkDa Cyt Tot P2 Mito

actin

SOD2

COX4

175836248

33

25

17

Page 58: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

58

4.3.2 Functional integry of the isolated mitochondria

Mitochondrial membrane potential ( ) was measured using JC-1 (5,5',6,6'-tetrachloro

1,1',3,3' tetraethylbenzimidazolylcarbocyanine iodide) dye. The dye undergoes a reversible

change in fluorescence emission from green to red as mitochondrial membrane potential

increases. JC-1 accumulates as aggregates in the mitochondria, resulting in red fluorescence.

The brightness of red fluorescence is proportional to . Succinate addition caused the

expected rise in , whereas uncoupling by CCCP (carbonyl cyanide 3-

chlorophenylhydrazone) led to depolarisation, indicating normal function of the mitochondrial

membrane.

Oxygen consumption was measured with an oxygraph on mitochondria respiring on

succinate. Addition of 2.5 M ADP causes increased oxygen consumption rate, coupled to

oxidative phosphorylation and yielding a mean respiratory control index (RCI) of 5.38,

indicating normal tightness (RCI between 3 and 10) of the coupling between oxidative

phosphorylation and respiration.

4.3.3 Isolation of mitoplasts

In order to obtain mitoplasts (mitochondria devoid of outer membrane), mitochondria were

incubated with a 5 × volume of cold hypotonic buffer (10 mM Tris, pH 7,4, 1 mM EDTA and

1 mM DTT ) for 10 min on ice. After 10 min, 150 mM NaCl was added and mitoplasts were

incubated 10 min on ice and centrifuged at 18000 g and at 4 C for 10 min. Mitoplasts were

resuspended into standard medium consisting of 0.3 M mannitol, 10 mM KCl, 10 mM

KH2PO4, 5 mM MgCl2, 1 mg/ml BSA, pH 7.4 to a concentration of 10 mg/ml.

4.3.4 Exposure of mitoplasts with cytosolic homogenates of G93A-SOD1 rat tissues

Five hundred micrograms of mitoplasts isolated from wt rat liver were exposed to 100 µg of

cytosolic spinal cord or cortex homogenate form 8 and 16 week old and end stage rats with

equalized amonts of human SOD1 at RT for 1 h. After the exposure, the mitoplasts were

centrifuged at 10000 × g for 1 min and rinsed three times with standard medium to wash away

unbound SOD1. An aliquot of each sample was resuspended in 1 × Laemmli sample buffer

and analyzed with anti-SOD1 Western blotting for the bound SOD1. The remaining mitoplast

suspension was analyzed for ROS production.

Page 59: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

59

4.3.5 Measurement of ROS production

ROS measurements were carried out in 96-well plates by mixing 200 µg (20 µl) of mitoplast

or mitochondria suspension with 140 µl standard medium with final concentrations of 1.3 mM

succinate, 3 µM antimycin, 10 µM 2,7-dichlorodihydrofluorescein diacetate (DCF, Sigma)

and 10 µM cytochrome c. ROS production was measured as fluorescence of oxidized DCF for

2h with Wallac Victor2 1420 multilabel reader (PerkinElmer, Waltham, MA, USA).

4.3.6 Isolation of intermembrane space and measurement of SOD1 activity

In order to isolate the contents of the intermembrane space, mitochondria (10 mg/ml) were

treated with 0.1 mg of digitonin per mg of mitochondria for 1 h at room temperature.

Iodoacetamide was added to samples to a concentration of 100 mM in order to prevent SOD1

activation upon disruption of the outer membrane. After centrifugation at 10 000 g and at

4 C for 10 min, the supernatant with intermembrane space was collected and saved. SOD

activity was assayed as quenching of NBT reduction by xantine oxidase/xanthine reaction

generated superoxide anion radical (Beauchamp and Fridovich, 1971; Oberley and Spitz,

1984). The optical density was measured with Wallac Victor2 1420 multilabel reader

(PerkinElmer). SOD1 activity was measured in the mitocondrial intermembrane space

preparation, obtained in presence of iodoacetamide, and was expressed as percentage of

activated enzyme in the absence of iodoacetamide (Inarrea et al., 2005).

4.3.7 SOD activity with zymography

SOD activity in the intermembrane space preparations was assessed as described (Beauchamp

and Fridovich, 1971). Two micrograms of total protein was loaded onto 10% native

acrylamide gels. After electrophoresis, the gels were washed in 50 mM phosphate buffer pH

7.8 for 10 min, then incubated in 1 mg/ml NBT solution in the same buffer for 15 min. After

incubation, the gels were briefly washed in phosphate buffer and incubated for 15 min in

0.25% TEMED solution containing 30 µM riboflavin. The gels were rinsed in phosphate

buffer and illuminated for 15 min with a fluorescent light source. SOD activity appears as

clear bands on a blue background. The gels were scanned with GS-710 Densitometer (Bio-

Rad) scanner and bands quantified with ImageQuant software (Molecular Dynamics,

Sunnyvale, CA, USA).

Page 60: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

60

4.4 Western blotting (I-III)

4.4.1 Sample preparation

Spinal cord, brain and liver from saline perfused animals were homogenized in 20mM sodium

phosphate buffer (pH 6.5) containing 0.225 mM PMSF, 1 mM EDTA, 0.5 g/ml leupeptin

and 0.5 g/ml aprotinin. Homogenates were centrifuged at 13,000 g for 15 min and

supernatants or the cytosolic fractions with soluble proteins were collected. The protein

concentration was measured with Bio-Rad Protein assay dye (Bio-Rad). For electrophoresis,

sampes were mixed with an equal volume of 2 Laemmli sample buffer, boiled for 5 min,

centrifuged at 13000 rpm for 30 s and loaded on to gels.

4.4.2 Electrophoresis and transfer

Polyacrylamide gel electrophoresis with SDS (SDS-PAGE) was carried out on Mini-Protean

3 (Bio-Rad) device at 200V constant voltage according to manufacturer's recommendations

on 10% or 12% polyacrylamide gels. After electrophoresis, the proteins were transferred on to

Hybond P polyvinylidene fluoride membrane (GE Healthcare, Uppsala, Sweden) in Mini

TransBlot (Bio-Rad) chamber. Membranes were blocked with 5% skimmed milk solution in

phosphate buffered saline containging 0.2% Tween and incubated with primary antibodies.

In addition to SDS-PAGE, the stability of SOD1 was analyzed using non-reducing

PAGE and native gradient PAGE. In non-reducing PAGE, the samples were not boiled and

the sample buffer did not contain -mercaptoethanol, leaving intramolecular disulphide bonds

intact. In native gradient PAGE, no SDS or other detergents were present and proteins were

left in native conformations for Western blotting with anti-SOD1 antibody. The native PAGE

gradient gels had a concentration gradient from 4 to 16% of acrylamide and were prepared

with gradient mixer (Sigma). Native sample buffers did not contain any detergents or

reducing agents, nor were the samples boiled for native PAGE.

4.4.3 malPEG modification of free cysteines

For the analysis of free cysteines, proteins were modified with 3 mM malPEG (mono-Methyl

polyethylene glycol 5'000 2-maleimidoethyl ether, Sigma) at 25 C for 1 h as described (Ferri

et al., 2006). The addition of malPEG to accessible free cysteines increases the mass of SOD1

by 5 kDa per modification and the shift in the molecular weight can be detected by anti-SOD1

immunoblotting following SDS-PAGE.

Page 61: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

61

4.4.4 Antibodies

The following primary antibodies were used in this study: Monoclonal anti- -actin (dilution

1:4000, Sigma, St. Louis, MO) was used as a loading control and a marker for the cytosolic

cellular fraction. COX4 (mouse monoclonal, 1:1000 dilution, Molecular Probes, Eugene, OR,

USA) and SOD2 (rabbit polyclonal, 1:5000 dilution, Stressgen Bioreagents, Ann Arbor, MI,

USA) were used as mitochondrial markers. Proteasome functions were studied using rabbit

polyclonal proteasome 20S LMP7 (dilution 1:1000, Abcam, Cambridge, UK), anti-

proteasome 20S X (dilution 1:1000, Abcam) and anti-ubiquitin antibody (dilution 1:1000,

Dako, Glostrup, Denmark). In addition, rabbit polyclonal anti-GLT-1 (dilution 1:1000,

Calbiochem, LaJolla, CA, USA), anti-PDI (rabbit polyclonal, 1:4000 dilution, Stressgen

Bioreagents) and anti-SOD1 (rabbit polyclonal, 1:4000 dilution, Stressgen Bioreagents) were

used. Anti-DNP antibody (anti-DNP, rabbit polyclonal, dilution 1:1000, Dako) was utilized to

detect DNPH derivatized oxidized proteins.

The following secondary antibodies were used for detection: HRP-labeled anti-mouse

IgG (dilution 1:4000, Amersham Biosciences, Uppsala, Sweden) or HRP-labeled anti-rabbit

IgG (dilution 1:3000, Amersham Biosciences), followed by ECL+ -detection (Amersham

Biosciences). Cy5-conjugated anti-mouse or anti-rabbit IgG (dilution 1:800, Jackson Immuno

Research, West Grove, PA, USA). Quantifications were done on a STORM Imager

(Molecular Dynamics, Sunnyvale, CA) with ImageQuant software (Molecular Dynamics).

4.5 Immunohistochemistry (I)

Three coronal 20 m cryosections from two animals per group of the lumbar spinal cord were

used for immunohistochemical labelling with rabbit polyclonal anti-proteasome 20S LMP7

(dilution 1:500, Abcam, Cambridge, UK) or rabbit polyclonal anti proteasome 20S X

antibodies (dilution 1:500, Abcam) and with cell markers monoclonal GFAP (glial fibrillary

acidic protein, dilution 1:500, Chemicon, Temecula, CA, USA), monoclonal NeuN (dilution

1:500, Chemicon) or monoclonal CD68 (dilution 1:500, Serotec, Oxford, UK) antibodies. For

the detection of proteasome immunoreactivities, Alexa Fluor 488 conjugated anti-rabbit IgG

antibody (Molecular Probes, Sunnyvale, CA) was used and for the detection of cell markers,

Alexa Fluor 568 conjugated anti-mouse IgG antibody (Molecular Probes) was used.

Fluorescence stains were visualized using a confocal microscope (Bio-Rad Radiance Laser

Scanning Systems 2100, Bio-Rad Microscience Ltd., Hertfordshire, UK) with a Laser-Sharp

2000 software (Bio-Rad Microscience Ltd.). For the visualization of proteasome

Page 62: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

62

immunoreactivity in light microscopy (Olympus AX70, Tokyo, Japan), incubation with

biotinylated anti-rabbit IgG (Vector Laboratories Inc., Burlingame, CA, USA) antibody was

followed by incubation with avidin–biotin complex (Vectastain Elite kit, Vector Laboratories

Inc., Burlingame, CA, USA) before the immunoreactions were visualized using nickel

enhanced diaminobenzidine (Sigma, St. Louis, MO) as a substrate.

4.6 Electrophoretic mobility shift assay (I)

Electrophoretic mobility shift assays (EMSA) for NF- B - DNA binding were carried out as

described earlier in detail (Helenius et al., 1996) with 5 µg of nuclear protein of the spinal

cord tissue at the presymptomatic age (100 d) and end-stage of G93A rats. Double-stranded

oligonucleotides for NF- B binding sites were from Santa Cruz Biotechnology (Santa Cruz

Biotechnology, Santa Cruz, CA, USA). The probe was labeled with T4 polynucleotide kinase

(Promega, Madison, WI, USA). Non-specific binding was blocked with 2 µg poly(dI-

dC) : poly(dI-dC) (Roche Applied Science, Basel, Switzerland) in a 20 µl assay volume.

Bound and free probes were separated in a native 4% polyacrylamide gel. Radioactive bands

were visualized with a STORM 860 Imager (Molecular Dynamics, Sunnyvale, CA) and pixel

volumes of specific bands were calculated with ImageQuant software (Molecular Dynamics).

4.7 Atomic absorption spectrophotometry (I)

Copper concentrations in the spinal cord, cortex and liver tissues from each treatment group

were measured by atomic absorption spectrophotometry at the City of Kuopio Environmental

Health Laboratory by Hitachi Z-8100 Polarized Zeeman (Hitachi, Tokyo, Japan) graphite

furnace atomic absorption spectrofotometry from pyrolyzed samples. Copper concentrations

were shown as mg of copper per g of tissue wet weight.

4.8 Proteasomal activity (I)

Proteasomal activity was measured from cytosolic fractions of the spinal cord samples as

chrymotrypsin-like activity by cleavage of suc-LLVY-amc (Sigma). The activity was

measured as increasing fluorescence of cleaved amc-peptides. Tissues were homogenized in

ice-cold buffer (50 mM Tris-HCl pH 7.5, 1 mM DTT, 0.25 M sucrose, 5 mM MgCl2, 0.5 mM

Page 63: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

63

EDTA, 2 mM ATP) and centrifuged at 1200 g and at 4 C for 7 min. The protein

concentration of the supernatant was determined using protein assay reagent (Bio-Rad).

Proteasomal activity was measured from aliquots of 10 g of protein in 50 l volume of assay

buffer; 20 mM Tris-HCl pH 7.5, 1 mM ATP, 2 mM MgCl2 and 0.1% BSA containing 100 M

suc-LLVY-amc. Fluorescence of cleaved amc-peptides was monitored every 10 min at 37 C

at 355 nm excitation and 460 nm emission using a Wallac Victor2 1420 multilabel counter

(PerkinElmer, Waltham, MA, USA).

4.9 Proteomics (IV)

4.9.1 Protein extraction

Samples from tg G93A-SOD1 mice, corresponding non-transgenic control mice and mice

expressing human wt SOD1 were kindly provided by Dr. Catherine Bendotti. Cytosolic

proteins were extracted from the spinal cords of 8, 14 and 18-week-old tg mice expressing

human G93A-SOD1, from the cerebella and hippocampi of 18-week-old tg mice and from the

corresponding non-tg control groups. In addition, proteins were extracted from the spinal

cords of 12-week-old tg mice expressing human wt SOD1. Tissues were homogenized with

Pellet Pestle Motor -homogenizator (Kontes, Vineland, New Jersey, USA) in 20 mM sodium

phosphate buffer (pH 6.5) containing 0.225 mM PMSF, 1 mM EDTA, 0.5 g/ml leupeptin

and 0.5 g/ml aprotinin. Homogenates were centrifuged at 13,000 g for 15 min and the

supernatant or the cytosolic fraction with soluble proteins was collected for analysis. The

protein concentrations were measured using Bio-Rad Protein Assay Dye (Bio-Rad, Hercules,

CA, USA) according to the manufacturer's instructions.

4.9.2 Carbonyl derivatization and detection

Protein oxidation was analyzed by derivatizing carbonyl groups, which are a hallmark for

oxidative damage, with 2,4-dinitrophenylhydrazine (DNPH). Twenty micrograms of protein

was derivatized with 2 mM DNPH for 1 h at room temperature. Proteins were precipitated

with 10 % TCA and washed three times with 1:1 ethanol:ethyl acetate. Samples were

dissolved back to 2D-PAGE sample buffer (9.5 M urea; 2 % triton X-100; 5 % -

merkaptoethanol; 1.2 % Bio-Lyte 5/7 ampholyte; 0.8 % Bio-Lyte 3/10 ampholyte) or

Laemmli sample buffer (62.5 mM Tris-HCl, pH 6.8; 2 % SDS; 5 % -mercaptoethanol; 25 %

Page 64: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

64

glycerol; 0,01 % bromophenolblue). The oxidized proteins were visualized by

immunoblotting and immunostained by anti-DNP antibody (anti-DNP, rabbit polyclonal,

dilution 1:1000, Dako) and secondary Cy5-labelled anti-rabbit IgG (Jackson Immuno

Research, West Grove, PA, USA). The fluorescence was detected with a STORM

fluoroimager (Molecular Dynamics). The relative degree of oxidation was measured as the

ratio of protein bound carbonyl immunoreactivity per total protein.

4.9.3 Two dimensional electrophoresis

Two dimentional (2D) electrophoresis was performed using Mini Protean 2D Cell (Bio-Rad,

Hercules, CA, USA) according to established procedures by Bio-Rad using isoelectric

focusing (IEF) in the first dimension and SDS-PAGE as the second dimension. Briefly, IEF

was performed with 4 % capillary acrylamide gels, containing 9.2 M urea; 20 % triton X-100;

1,2 % Bio-Lyte 5/7 ampholyte; 0.8 % Bio-Lyte 3/10 ampholyte with 1000V for 5 h. After the

first dimension, IEF gels were reduced with buffer containing 2.3 % SDS; 62,5 mM Tris-HCl,

pH 6.8; 5 % - mercaptoethanol; 10 % glycerol; 0,01 % bromophenolblue and the second

dimension was performed with 12% SDS-PAGE. After electrophoresis, the proteins were

transferred to PVDF membrane and total proteins were stained using fluorescent Sypro Ruby

stain (Bio-Rad). The fluorescence was detected with a STORM fluoroimager (Molecular

Dynamics) and proteins were quantified using ImageMaster software (Amersham

biosciences).

4.9.4 In-gel digestion

Protein spots of interest were excised from Sypro Ruby (Bio-Rad) stained gels and digested

with trypsin according to UCSF in-gel digestion protocol

(http://donatello.ucsf.edu/ingel.html). Briefly, gel pieces were first minced and dehydrated by

washing with 25 mM NH44CO3 containing 50% acetonitrile. Dehydrated gel matrix was dried

with vacuum centrifugation (Jouan RC 10.10, Nantes, France) for 20 minutes and rehydrated

with 25 mM NH44CO3 containing 0.1 g/ l trypsin (Promega, Madison, WI, USA). The

rehydrated gel matrix was covered with 25 mM NH44CO3 and proteins were digested

overnight at 37 C. Peptides were eluted form gel matrix three times with 50% acetonitrile

containing 5% TFA and the eluate was concentrated back to a volume of 10 µl with vacuum

centrifugation.

Page 65: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

65

4.9.5 Mass spectrometry

Digested peptides were identified with liquid chromatography electro spray ionization tandem

mass spectrometry (LC-ESI-MS/MS). Peptides were separated with Ultimate/Famos LC

capillary liquid chromatography (LC Packings, Amsterdam, NL). Peptides were first purified

and concentrated on a 300 µm id. 1 mm C18 PepMap column for 7 minutes with 0.1 %

acetic acid on a 10µl/minute flow rate. Separation of peptides for MS was done on a 75 m

id. 50 mm analytical C18 PepMap column where the sample was eluted with 3-40%

acetonitrile gradient containing 0.1% acetic acid for 40 min with a flow rate of 200 nl/min.

LC was connected to the mass spectrometry by Protana electro spray platform (Protana,

Odens, Denmark) and 30µm id. PicoTip capillary (New Objective, Woburn, MA, USA).

Peptide mass and fargmentation spectra were measured on quadruple ion-trap mass

spectrometry (Finnigan MAT, San Jose, CA, USA) on 300-4000 mass/charge (m/z) range.

The fragmentation spectrum was obtained by collision induced dissociation with helium. The

data were analyzed on Xcalibur software with Sequest algorithm comparing the spectra on

Swiss-Prot and NCBI protein databases.

4.10 Flow cytometry (III)

4.10.1 Peroxide production in mouse blood lymphocytes

Mouse (wt and SOD1-/-) blood lymphocytes were isolated with differential centrifugation

using Ficoll Paque PLUS (GE Healthcare) according to the manufacturer's instructions.

Lymphocytes were washed and suspended in HBSS containing 10 µM DCF. Ten minutes

later 3 µM antimycin A was added and the samples were analyzed with FACSCalibur flow

cytometer (Becton-Dickinson, Franklin Lakes, NJ, USA) at the time poins of 0 min, 30 min

and 90 min. Altogether, 10000 cells per sample were counted.

4.10.2 Antimycin A-induced apoptosis in lymphocytes

Antimycin A-induced apoptosis was measured by AnnexinV-FITC binding in mouse (wt and

SOD1-/-) blood lymphocytes that were isolated with differential centrifugation in Ficoll Paque

PLUS (GE Healthcare) according to manufacturer's instructions. Lymphocytes were washed

and suspended in HBSS containing 20 µM antimycin A. Aliquots of suspensions containing

Page 66: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

66

non-treated cells or cells incubated with antimycin A for 30, 60 or 90 min were sedimented by

centrifugation and resuspended in binding buffer containing 10mM HEPES, 140mM NaCl,

2.5 mM CaCl2, pH 7.4. Annexin V-FITC (Sigma) was added to a final concentration of

0.45µg/ml. After a 10 min incubation, propidium iodide (PI) was added to a final

concentration of 2 µg/ml. After incubation for another 10 min, samples were analyzed with

FACSCalibur flow cytometer (Becton-Dickinson). Again, 10000 cells per sample were

counted altogether.

4.11 Isolation and purification of human SOD1 (II,III)

Human SOD1 was isolated from human erythrocytes and human G93A-SOD1 was isolated

from erythrocytes of G93A-SOD1 transgenic rats by fractionated precipitation and DEAE ion

exchange chromatography (Yoo et al., 1999). The purity and identity of isolated SOD1 was

analyzed by SDS-PAGE and anti-SOD1 immunoblotting.

4.12 Measurements of cytochrome c catalysed peroxidation (III)

To study the role of SOD1 in cytochrome c-catalysed DCF oxidation in the presence of a

superoxide source, 15 l of xanthine oxidase suspension (Sigma), diluted 1:100 in 50 mM

phosphate buffer (pH 7.8), was mixed with xanthine solution to a 0.4 mM final concentration

and complemented with 10 M DCF. The total volume of the reaction mixture was 200 l.

DCF oxidation was recorded on a Victor multilabel reader (Wallac). The fluorescence kinetics

were recorded in the presence of 10 M cytocrome c (Sigma) and together with purified

human erythrocyte SOD1 (15, 30, 60, 250 nM).

4.13 Statistical analysis (I-IV)

All results are shown as mean standard deviation. Statistical analyses were performed on

SPSS software version 11.5 (SPSS Inc Chicago, IL, USA). Two group comparisons were

done using Student’s t-test. For the comparison of three or more groups the statistical

signifigance was determined with one way ANOVA combined with Bonferroni post-hoc test.

p<0.05 was considered statistically significant.

Page 67: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

67

5. RESULTS

5.1 PDTC reduced survival of G93A-SOD1 ALS rats without affecting NF- B (I)

PDTC is considered to be a very potential drug candidate as it has been shown to be

protective in several models of CNS and peripheral diseases by inhibiting the activation of

transcription factor NF- B and serving as a strong antioxidant. NF- B activation may

promote expression of genes that mediate inflammation or apoptosis, as well as some genes

that support survival. Therefore, inhibition of NF- B has been considered to serve as a

beneficial target for drug treatments.

However, in a G93A-SOD1 rat model of ALS, PDTC treatment did not have a

beneficial effect, but in opposite, PDTC decreased the survival of G93A-SOD1 rats by 15%.

Moreover, the mechanism was not mediated through NF- as EMSA analysis of the spinal

cord samples showed no differences in DNA binding activity of NF- B between PDTC and

untreated groups.The survival ages for the groups were 140 13 days in the untreated group

and 122 21 days in the PDTC group (p<0.01). Compared to the untreated group of G93A-

SOD1 rats, also the onset of paralysis occurred significantly earlier in PDTC treated animals

(109 22 days and 120 14 days for PDTC and untreated, respectively, p<0.01), whereas

there was no difference in the duration of the disease from the onset to the end stage between

PDTC (12 4 days) and vehicle treated (11 3 days) ALS rats. Weight gain measured at or

after the onset of the disease showed no significant differences between PDTC and untreated

tg or wt rats and no significant signs of toxicity of the PDTC treatment were observed, as

judged by weight gain, consumption of drinking water, development of diarrhoea, copper

concentration of the liver, ataxia, overall locomotor activity, or immunohistochemical signs of

increased gliosis or myelin loss in the ventral spinal cord.

The result of reduced survival was truly unexpected, and moreover, the mechanism

was clearly independent of NF- B in opposite to our initial hypothesis. The results left us no

choice but to research the literature to find other possible targets for PDTC. The next possible

targets of PDTC studied were tissue copper concentrations, as PDTC is also a metal chelator

for transitional metals, and proteasome activity, as copper-PDTC complex may inhibit

preoteasomal activity.

Page 68: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

68

5.2 Copper levels were increased in ALS rat tissues and were further increased in spinal

cords by PDTC treatment (I)

As Cu,Zn-SOD1 contributes to the cellular copper concentration, copper concentrations of

tissues in G93A-SOD1 transgenic rats were significantly higher than in corresponding wild

type littermate rates due to SOD1 over-expression. The copper concentrations of spinal cord

were significantly higher in both presymptomatic (3.3 0.2 g/g, n=5, p<0.001) and end-

stage (3.1 0.5 g/g, n=7, p<0.001) G93A-SOD1 rats than in wt rats (1.7 0.1 g/g, n=4). In

addition to the increase caused by SOD1 over-expression, PDTC treatment increased the

copper levels of the spinal cord tissue in G93A-SOD1 rats by 36% (4.2 0.8 g/g, n=7,

p<0.05) and in wt rats by 200% (5.1 1.9 g/g, n=4, p<0.001). There was no statistically

significant difference in the spinal cord copper concentration between G93A-SOD1 and wt

rats after PDTC treatment. PDTC treatment did not cause significant increases in copper

concentration in the cortex or liver of G93A-SOD1 rats.

5.3 PDTC inhibited immunoproteasome (I)

As Cu-PDTC complexes may interfere with the proteasome, we investigated whether

proteasomal activity is changed by PDTC treatment. Proteasome activity was measured as

chymotrypsin-like activity in the spinal cord tissues. Chymotrypsin-like activity was

approximately on the same level in the untreated wt group and the untreated G93A-SOD1

group at the presymptomatic stage. In the end stage, the G93A-SOD1 rats had significantly

increased proteasomal activity and this increase was completely prevented by PDTC

treatment. PDTC treatment had no effect on proteasome activity in wt animals. The results on

proteasome activity were further supported by the notion that levels of ubiquitinated proteins

in the spinal cord of PDTC treated tg rats had increased significantly. Immunoblotting

revealed that the levels of ubiquitinated proteins were increased by 33% in the cytosolic

fraction of the spinal cord of PDTC treated tg rats at the end stage of the disease when

compared to untreated end stage animals.

To investigate in detail whether PDTC alters the expression of specific proteasome

subunits in the spinal cord, we used immunoblotting for 20S X and 20S LMP7, markers of

constitutive and inducible proteasome, respectively. The level of 20S X was not changed in

G93A-SOD1 rats at any time point of the disease progression. Instead, the expression of

immunoproteasome measured by immunoblotting for 20S LMP7, an inducible subunit, was

strongly increased in the spinal cord but not in the cortex along with the disease progression

Page 69: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

69

of G93A-SOD1 rats. The amount of LMP7 protein was increased 6-fold between 8 and 16

weeks of age (p<0.01, when 8w and 16w compared, respectively), and reached a 9-fold

increase at the end stage (p<0,001). PDTC treatment completely prevented the induction of

20S LMP7 at the end stage of G93A-SOD1 rats (p<0.05), whereas no effect of PDTC on the

constitutive proteasome subunit was detected. Importantly, in wt animals, 20S LMP7 was

barely detectable or undetectable in the cytosolic fraction, and PDTC had no effect on

expression of this protein, or 20S X in wt animals.

5.4 PDTC upregulated GLT-1 (I)

Although PDTC has also anti-inflammatory effects, the changes in immunoproteasome levels

after PDTC treatment were not due to a common reduction of astroglial functions, as PDTC

also increased the levels of astrocyte specific glutamate transporter (GLT-1), a potential drug

target in brain diseases. In the spinal cords of untreated tg rats the levels of GLT-1 were

decreased, whereas in PDTC treated tg rats the levels of GLT-1 were at the same levels as in

wt rats.

5.5 PDTC prevented glial immunoproteasome induction (I)

Initial light microscopy imaging of the lumbar spinal cord sections for constitutive (20S X)

and immunoproteasome (20S LMP7) subunits showed that the constitutive proteasome was

expressed in cells throughout the gray matter in the ventral horn of the lumbar spinal cord in

both PDTC treated tg and wt rats, and in untreated tg and wt rats. However, the

immunoproteasome was expressed only in untreated tg rats at the end stage. From the

appearance, the cells showing immunoproteasome staining looked non-neuronal. Double

labeling immunohistochemistry with confocal imaging for cell markers for neurons (NeuN),

astrocytes (GFAP) and microglia (CD68), and for constitutive proteasome (20S X) and

immunoproteasome (20S LMP7), showed that the immunoproteasome 20S LMP7 was

expressed in astrocytes and microglia, whereas proteasome 20S X was also expressed in

neurons. The results indicate that immunoproteasome indution occurs at the end stage of the

disease in microglia and astrocytes, and that PDTC is able to inhibit glial immunoproteasome

induction with devastating effects on survival of the rats.

Page 70: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

70

5.6 Mutant SOD1 was oxidized and destabilized in spinal cords of G93A-SOD1 rodent

models (II,IV)

Proteomic analysis of protein expression and oxidation during pathogenesis in G93A-SOD1

mice from presymptomatic, symptomatic and end stage of the disease revealed that mutant

SOD1 is one of the oxidized proteins in the spinal cord of transgenic G93A-SOD1 mouse

model of ALS. Moreover, the oxidation of G93A-SOD1 increases significantly with disease

progression specifically in spinal cord as the level of oxidation of G93A-SOD1 was higher in

spinal cord when compared to oxidation in less affected regions of the CNS, such as the

cerebellum and hippocampus. In addition, with two-dimensional electrophoresis followed by

mass spectrometric identification we detected a fragment of human SOD1 that was oxidized.

Moreover, the mouse endogenous SOD1 was not oxidized at all, or its oxidized form is

promptly degraded, implicating a possible toxic property for the oxidized form. However, also

the human wt SOD1 is oxidized to the same extent as the as the G93A-SOD1 in the spinal

cord as shown by 2D anti-DNP immunoblotting of spinal cord from human wt SOD1

expressing mice. Analysis of protein expression also revealed that expression of PDI, a

molecular chaperone which rearranges inter- and intramolecular disulphide bonds between

thiol groups of cysteine residues on unfolded proteins (Turano et al., 2002), was upregulated

early in the disease progression, before any symptoms were seen. This suggest that thiol-

related oxidation is altered in ALS, and that the status of the disulphide bonds in SOD1 may

greatly affect the stability of the SOD1 protein.

The stability of SOD1 was analyzed by employing modified immunoblotting, which

allows distinguishing the degree of denaturation and loss of quaternary structure by binding to

a hydrophobic PVDF membrane in non-reducing conditions. Native SOD1 is an exceptionally

stabile protein dimer and in non-reducing electrophoretic conditions SOD1 should appear as a

dimer. The non-reducing PAGE analysis of purified G93A-SOD1, SOD1 from the spinal

cord, cortex and liver samples of G93A-SOD1 rats revealed that purified G93A-SOD1 and

G93A-SOD1 from the liver were present as a dimer, indicating that protein stability of the

purified mutant SOD1 or mutant SOD1 in peripheral tissues, as assessed by electrophoretical

behavior, is not altered. However, SOD1 seemed to be less stable in the cortex and

cerebellum, since the SOD1 dimer from the cortex and cerebellum had partially broken down

to monomers. In the spinal cord, the quaternary structure of SOD1 was destabilized the most

as SOD1 in the spinal cord had less of both dimeric and monomeric conformations compared

to the cortex, and the majority of SOD1 in the spinal cord was present as a misfolded

monomer having a molecular weight of approximately 20-30 kDa. The monomeric

Page 71: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

71

destabilized SOD1 with a molecular weight of 20-30 kDa was present in the spinal cord

throughout the disease from the early presymptomatic time point of 8-weeks until the end

stage of the disease, whereas non-reducing PAGE analysis of SOD1 from the cortex,

crebellum or liver showed no destabilization or misfolded monomers with disease

progression.

Surprisingly, purified human wt SOD1 did not bind to Western blotting membrane in

the non-reducing conditions at all, possibly due to the fact that native wt SOD1 is an

exceptionally stable and hydrophilic protein dimer and therefore may not bind to the

hydrophobic blotting membrane under the non-reducing conditions. It is also important to

note that the differences seen in electrophoretical mobility of mutant SOD1 from various

tissues are not due to incorporation of insoluble aggregates to the sample, as native gradient

PAGE did not show any high molecular weight bands for the samples. This indicates that

aggregation and misfolded monomeric species are originating from the soluble biologically

stable dimeric SOD1 detected in the native immunoblot.

5.7 PDI was upregulated and its levels inversely correlated with the levels of cysteine

reduced SOD1 (II,IV)

As PDI was upregulated early in the disease progression as shown by proteomic analysis, and

as mutant SOD1 was clearly destabilized over disease progression in affected tissues, the

nature of the destabilization was further investigated by malPEG assay, where the oxidation

state of the SOD1 cysteine residues were analyzed using malPEG modification and anti-

SOD1 Western blotting. Human SOD1 has four cysteines and only the reduced and exposed

cysteine residues are expected to react with malPEG. The cytosolic spinal cord and cortex

samples were modified with 3 mM malPEG, which will react and bind to free reduced

cysteine residues, causing a 5 kDa increase in molecular weight that can be detected in anti-

SOD1 Western blot. MalPEG modification showed a significant increase in the levels of free

reduced cysteine residues of SOD1 in the spinal cord at the end stage of the disease when

compared to the cortex. In the spinal cord, malPEG modification produced size shifts of both

5 kDa and 10 kDa, indicating a modification of one and two cysteine residues, respectively.

As levels of the disulphide-reduced SOD1 had indeed increased in tg mice, we

confirmed our results of the proteomics data, showing increased PDI expression in G93A-

SOD1 mice, also in tg rats with anti-PDI western blotting over disease course in G93A-SOD1

rats. PDI is a molecular chaperone, which functions to rearrange inter- and intramolecular

Page 72: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

72

disulphide bonds between cysteine residues of unfolded proteins (Turano et al., 2002). Human

SOD1 has four cysteine residues (C6, C57, C111 and C146). Two of these, C57 and C146,

form an intrasubunit disulphide bond. Expression of PDI was upregulated before the disease

onset at 8 weeks and at the end stage was restored back to levels similar to the cortex of

mutant SOD1 expressing rats or the spinal cord of wt rats. These results are in line with the

proteomics data showing PDI upregulation at presymptomatic stage in G93A-SOD1 mice.

PDI upregulation in ALS models has also been shown by others (Atkin et al., 2006). These

data suggest that early upregulation of PDI in the disease progression may be an attempt to

cope with accumulation of destabilized SOD1 but is eventually overwhelmed by aggregation

of destabilized SOD1 with disease progression.

5.8 Mitochondrial SOD1 levels were the highest in the spinal cord of G93A rats (II)

Mitochondria are a likely target of mutant SOD1 toxicity and as SOD1 is also expressed in

the intermembrane space of mitochondria, we analyzed the levels of mutant SOD1 localized

to mitochondria from isolated and purified mitochondria with anti-SOD1 Western blotting.

Mitochondria were isolated from transgenic G93A-SOD1 rat cortex and spinal cord tissues

form different stages of the disease: 8-weeks - presymptomatic, 16-weeks - onset, and from

the end stage. The purity of the mitochondrial fraction was determined with Western blotting

as a presence of mitochondrial markers COX4, SOD2 and absence of cytosolic stuctural

component actin. SOD1 amounts were quantified from anti-SOD1 immunoblots and

normalized against COX4. SOD1 levels in mitochondria were 40-100% higher in the spinal

cord when compared to the cortex at presymptomatic stage and at the end stage of the disease.

5.9 Mutant SOD1 bound to mitoplasts and enhanced ROS production (II)

As the association of mutant SOD1 to mitochondria may also be related to the over-

expression of the transgene (Bergemalm et al., 2006), we set up a model system for mutant

SOD1 toxicity in mitochondria where liver mitoplasts (mitochondria devoid of outer

membrane) from wt rats were exposed to cytosolic tissue homogenates of G93A-SOD1 rats

from the spinal cord and cortex, with or without destabilized SOD1, respectively. The purity

of the mitoplast fraction was determined with Western blotting as a presence of mitochondrial

marker SOD2 and absence of cytosolic stuctural component actin. In addition, rat endogenous

Page 73: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

73

SOD1 resided in the mitoplast fraction. The binding of destabilized mutant SOD1 from spinal

cord cytosol to the mitoplasts, isolated from wt rat liver, was increased after exposure when

compared to binding of SOD1 deriving from the cortex. In parallel, ROS production was

significantly elevated in mitoplasts exposed to the cytosolic spinal cord homogenate of an 8-

week-old G93A-SOD1 rat.

5.10 Mutant SOD1 increased hydrogen peroxide production in the intermembrane space

of mitochondria (III)

The role of SOD1 in mitochondria is somewhat controversial as dismutase activity is not

required in the intermembrane space. Instead of SOD1, cytochrome c can quench superoxide

efficiently in the intermembrane space by oxidizing superoxide directly to oxygen. Therefore,

cytochrome c acts as a true antioxidant (Pereverzev et al., 2003), whereas SOD1 will produce

hydrogen peroxide in addition to oxygen. We hypothesized that upon mitochondrial stress

SOD1 might compete with cytochrome c for superoxide in the intermembrane space and

generate hydrogen peroxide, which then could react with cytochrome c and oxidize the

cytochrome c molecule to oxoferryl heme. Oxoferryl heme is a highly reactive oxidant that is

able to react with a number of intracellular targets including proteins, nucleic acids and lipids

(Lawrence et al., 2003), eventually leading to a paradoxical increase in ROS production and

cellular injury.

The mechanism for increased mitochondrial ROS production and the role of SOD1

was further studied by isolated wt liver mitochondria challenged with antimycin A, an

inhibitor of complex III, resulting in a break in the electron transport chain of the oxidative

phosphorylation and a prompt superoxide production as shown by electron paramagnetic

resonance (EPR) (Han et al., 2003). The SOD activity in an intermembrane space preparation

was rapidly increased as a function of time in response to antimycin A. Mitochondrial

respiration also resulted in increased production of hydrogen peroxide as determined in

parallel by measurement of the fluorescence of 2,7-dichlorodihydrofluorescein diacetate

(DCF), a widely used probe sensitive for hydrogen peroxide (Hempel et al., 1999). The

response remarkably coincided with the maximal increase in SOD activity. SOD1 was

essential for the increased hydrogen peroxide production, because adding SOD1 inhibitors

ammonium tetrathiomolybdate or ammonium diethyldithiocarbamate, significantly and dose-

dependently reduced the DCF fluorescence in isolated mitochondria. Moreover, the

mitochondria isolated from SOD1 deficient knock out mice (SOD1-/-) mice produced

Page 74: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

74

substantially less DCF fluorescence and did not show any response to inhibition of complex

III by antimycin A. Thus, the elevated SOD1 activity is responsible for the increased

hydrogen peroxide production in the intermembrane space, resulting in cytochrome c-

catalysed DCF oxidation.

In order to be sure that the seen effects were not due to damage caused to

mitochondria by isolation, the functional integrity of the mitochondria was ensured by

measuring membrane potential with JC-1 dye, and the respiration rate of the isolated

mitochondria was measured as oxygen consumption with an oxygraph. The membrane

potential increased with addition of succinate, allowing respiration, and decreased after

uncoupling with CCCP, which leads to depolarisation, indicating normal function of the

mitochondrial membrane. The respiration rate increased after addition of ADP and was

coupled to oxidative phosphorylation, indicating normally respiring mitochondria measured

as oxygen consumption with an oxygraph. Also L-NNA (N -Nitro-L-arginine), an inhibitor of

nitric oxide synthesis (NOS), had no effect on the hydrogen peroxide production, indicating

that the peroxide measure with DCF is not preoxynitrite (-OONO) formed by reaction with

superoxide and NO produced by the NOS.

In order to further test the hypothesis that SOD1 activity is responsible for the

increased hydrogen peroxide production in the intermembrane space, we isolated mitoplasts,

i.e. mitochondria devoid of outer membrane and the intermembrane space, to reconstitute

conditions for hydrogen peroxide production. No DCF oxidation could be detected in

respiring mitoplasts even after inhibiting complex III by antimycin A, indicating that the auto-

oxidation rate of DCF without peroxidase is low. Addition of cytochrome c caused a slight

increase in DCF fluorescence, possibly due to hydrogen peroxide escaping from the

mitochondrial matrix. Importantly, addition of SOD1 at 100 nM concentration more than

doubled the rate of DCF oxidation in the presence of cytochrome c. To confirm the role of

peroxide in the DCF oxidation, addition of horseradish peroxidase (HRP) to mitoplasts was

shown also to increase the DCF oxidation. Addition of SOD1 together with HRP increased

the oxidation even more, demonstrating also the contribution of superoxide dismutation.

To model the interaction of superoxide, cytochrome c and SOD1 in the intermembrane

space, we reconstituted a reaction where superoxide was generated through xantine

oxidase/xantine (XO/X). A slow oxidation of DCF occurred in the presence of this enzyme

substrate pair. The rate of DCF oxidation was slightly elevated by 5 M cytochrome c.

However, adding increasing concentrations of SOD1 in the reaction mixture strongly and

dose-dependently increased the rate of DCF oxidation, indicating that SOD1 significantly

enhances cytochrome c-catalyzed peroxidation.

Page 75: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

75

To investigate whether SOD1 controls the production of hydrogen peroxide in the

mitochondrial intermembrane space also in intact cells, we isolated lymphocytes from wt and

SOD1-/- mouse blood by differential centrifugation, and loaded them with DCF before adding

antimycin A. Flow cytometric analysis showed that antimycin A-induced DCF oxidation was

attenuated in SOD1-/- lymphocytes at 90 min of incubation as compared to the hydrogen

peroxide production in wt lymphocytes. In addition to increased superoxide production in

mitochondria, inhibition of complex III by antimycin A has also been shown to induce

apoptosis (Wolvetang et al., 1994). To test whether SOD1 activity is important for apoptosis,

SOD1-/- and wt lymphocytes were challenged with antimycin A for 90 min and apotosis was

determined by Annexin V binding using flow cytometry. SOD1-/- lymphocytes showed

significantly less apoptosis than wt lymphocytes.

5.11 Mutant SOD1 activity and ROS production were increased in spinal cord

mitochondria of G93A-SOD1 mice (III)

According to these data, the increased amount of SOD1 associated with the mitochondrial

intermembrane space would be expected to cause enhanced ROS production. We studied

mitochondrial SOD1 activity in mutant G93A-SOD1 mice, and in fact, the SOD1 activity of

intermembrane space preparations from the spinal cord was 6-fold greater in G93A-SOD1

mice than in wt mice. In addition, SOD1 activity in the mitochondrial intermembrane space in

spinal cord was somewhat higher in comparison with non-affected brain tissues of the same

animals.

The SOD1 activity of the mitoplast preparations isolated form the spinal cord was

twice as high as from the brain of G93A-SOD1 mice, whereas the SOD2 activity was equal in

mitoplasts between spinal cord and brain. To investigate whether the elevated activity and

accumulation of mutant G93A-SOD1 in mitoplasts results in increased ROS production, we

measured DCF oxidation in the mitoplast preparations in the presence of cytochrome c. The

mitoplasts isolated from the spinal cord of G93A-SOD1 mice produced significantly more

DCF fluorescence compared to the mitoplasts that derived from the cortex of G93A-SOD1

mice, indicating that the increased SOD1 amount in the intermembrane space is also

associated with an increase in hydrogen peroxide production. The main findings reported in

the results are summarized in table 2.

Page 76: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

76

Table 2. Main findings of the thesis.

Originalpublication Main findings

Referenceto chapter

IPDTC reduced survival of ALS rats by inhibiting induction ofALS pathogenesis specific immunoproteasome in glial cells andby increasing copper concentration in the CNS.

5.1 – 5.25.4

IImmunoproteasome inhibition lead to faster disease progressiondespite PDTC upregulated levels of glutamate transporter GLT-1 in the CNS of ALS rats.

5.3

II, IVMutant SOD1 was oxidized and destabilized specifically inspinal cord of ALS rats and mice as shown by carbonyl levelsand stability of SOD1 dimer to mild denaturation.

5.5

II, IVPDI was upregulated in spinal cord and levels of PDI inverselycorrelated with reduced disulphide bonds of mutant SOD1. 5.6

IIMutant destabilized SOD1 from spinal cord associated tomitoplasts (mitochondria devoid of outer membrane) andincreased ROS production.

5.7 – 5.8

IIIIncreased ROS production may be due to elevated SOD1activity in the intermembrane space of mitochondria leading toincreased hydrogen peroxide production and oxidation ofcytochrome c to oxoferryl heme.

5.9 – 5.10

Page 77: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

77

6. DISCUSSION

This study was carried out to investigate the stability and oxidation of human mutant SOD1

through the disease progression using tg G93A-SOD1 ALS rats and mice and to analyze the

role of mutant SOD1 in mitochondria. In addition, as inflammation and oxidative damage are

implicated greatly in ALS pathogenesis, we treated G93A-SOD1 rats with PDTC, a drug

known for its anti-inflammatory and antioxidative properties.

6.1 PDTC inhibits immunoproteasome induction resulting in reduced survival of ALS

rats (I)

Previous studies have shown that PDTC provides protection in a number of CNS and

peripheral disease models by inhibiting activation of transcription factor NF- B, serving as a

strong antioxidant, or by activating the protective Akt-GSK3 pathway (Cheng et al., 2006;

Cuzzocrea et al., 2002; La Rosa et al., 2004; Matsui et al., 2005; Messina et al., 2006; Nurmi

et al., 2006; Nurmi et al., 2004a). Even though several antioxidants (Gurney et al., 1996; Wu

et al., 2003) and inhibitors of NF- B-driven genes, such as COX-2, TNF and IL-1

(Drachman et al., 2002; Kiaei et al., 2006) prolong survival of tg ALS mice, and activation of

the Akt-GSK3 pathway reduced mutant SOD1-mediated motor neuron cell death in vitro

(Koh et al., 2005), we found that PDTC treatment does not provide protection, but instead,

significantly decreases the survival of G93A-SOD1 rats. The fact that PDTC treatment

prevented the reduction in levels of glutamate transporter GLT-1, a potential therapeutic

target verified in numerous animal studies of ALS (Gurney et al., 1996; Howland et al.,

2002), and that PDTC treatment did not induce toxic side effects in the tg or wt rats, the

negative finding is of interest. It implies that PDTC treatment caused some harmful alterations

in cellular functions which were able to override the potentially beneficial effects of the drug.

In agreement with the previous studies on mouse ALS models (Cheroni et al., 2005;

Puttaparthi and Elliott, 2005), we observed that both an increase in proteasome activity and an

induction of immunoproteasome selectively occur in the affected spinal cord tissue of a

G93A-SOD1 rat model. Importantly, we found that PDTC treatment strongly inhibited these

ALS-model specific changes in the proteasome. On the other hand, in line with previous

reports showing that PDTC is a metal chelator and transports copper from the extracellular

medium into the cell, we found that PDTC treatment increased the copper concentration in the

spinal cord of both G93A-SOD1 tg and wt rats. Considering that PDTC is also known as a

Page 78: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

78

proteasome inhibitor (Kim et al., 2004), and that increased copper concentration may be

needed for the proteasome inhibitory activity of PDTC, it is likely that the detrimental effect

of PDTC on G93A-SOD1 rats is mediated by inhibition of the immunoproteasome. This

inhibition of the immunoproteasome was also evidenced as increased levels of ubiquitinated

proteins in PDTC-treated G93A-SOD1 rats. Because 20S X, a marker of the constitutive

proteasome, was not affected in G93A-SOD1 tg rats alone or in G93A-SOD1 by PDTC

treatment, whereas 20S LMP7, an inducible subunit of immunoproteasome, was induced

selectively in astrocytes and microglia, our results suggest that the immunoproteasome in

non-neuronal cells plays a protective role in G93A-SOD1 rats.

In addition to its effect on proteasome acitivity, an increased copper concentration

may well trigger other cellular processes, which in animals overexpressing G93A-SOD1

could accelerate neurodegeneration. An oral, 15-day PDTC treatment at millimolar doses has

been reported to increase the copper concentration and levels of lipid peroxidation products in

rat peripheral nerves (Calviello et al., 2005), implicating that at least in some rodent tissues a

long-term PDTC treatment may cause oxidative stress associated with increased copper

concentrations. On the other hand, a 7-month PDTC treatment with the same protocol as in

the present study significantly increases cortical copper concentrations in a mouse model of

Alzheimer's disease and prevents the decline in cognition without increasing oxidative stress

in the brain (Malm et al., 2007). Even though induction of the immunoproteasome was

observed selectively in G93A-SOD1 rat spinal cords and PDTC prevented this induction

without causing changes in other known targets of PDTC, we cannot exclude the possibility

that the increased copper concentration enhances motor neuron degeneration in G93A-SOD1

rats also by other mechanisms in parallel with immunoproteasome inhibition. However, it

should be noted that the -lactam antibiotic ceftriaxone, which has been reported to be

neuroprotective in models of ALS by increasing expression of glutamate transporter GLT-1

(Rothstein et al., 2005), is also a metal chelator. PDTC resembles ceftriaxone as it

significantly increases expression of GLT-1 and copper concentration in the spinal cord. We

hypothesize that even though -lactams and PDTC might both be able to modulate GLT-1

and copper concentration, only PDTC but not -lactams also inhibits immunoproteasome. As

induction of immunoproteasome may be a rather selective characteristic for ALS (models)

compared to models of ischemia, trauma and amyloid accumulating diseases, inhibition of the

immunoproteasome alone in ALS models could increase the accumulation of ubiquitinated

proteins, including ubiquitinated SOD1, which has been suggested to gain neurotoxic

functions such as increased peroxidase activity.

Page 79: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

79

PDTC is an established inhibitor of NF- B. Even though inhibition of NF- B has

frequently been associated with tissue and cellular protection, inhibition of NF- B may also

accelerate neurodegeneration because of down-regulation of survival-supporting NF- B-

regulated genes, such as manganese SOD and Bcl-2 (Mattson and Camandola, 2001). Even

though we detected a trend for increased NF- B binding activity in the spinal cords of G93A-

SOD1 rats, no statistically significant differences between any of the untreated/PDTC treated

rats were observed. It is possible that in a chronic disease such as ALS, NF- B activity even

though being induced is not maintained at such high levels. On the other hand, oral PDTC

administration may not allow achieving as high tissue concentrations as intraperitoneal

administration does, and thereby does not result in efficient inhibition of NF- B. Also, we

cannot exclude the possibility that NF-kB binding to DNA is increased at time points other

than the presymptomatic (100 days) and end-stage time points studied here. Nevertheless, our

experiments do not provide evidence for a central role of NF- B in ALS. In a previous study,

NF- B immunoreactivity was found to be increased in astrocytes surrounding degenerating

motor neurons (Migheli et al., 1997). Therefore, the possible antiapoptotic effects of NF- B

may not be, at least, directly involved in neuronal survival in tg ALS models.

Nevertheless, PDTC inhibited the immunoproteasome induction and resulted in

decreased survival despite upregulated GLT-1 and other possible beneficial effects, including

anti-inflammatory and anti-oxidative mechanisms. These results indicate that

immunoproteasome induction in the disease pathogenesis may be an important mechanism for

coping with the toxic consequences of mutant SOD1 in tissues affected in ALS.

6.2 Mutant SOD1 oxidation and destabilization precede aggregation and loss of activity

(II, IV)

Mutant SOD1 is one of the oxidized proteins in the spinal cord of G93A-SOD1 mice and rats

and the level of oxidation increases with disease progression. However, while the mouse

endogenous SOD1 remains intact, the human mutant and wild-type SOD1 are oxidized to the

same extent in the spinal cord. Therefore, the oxidative response to the presence of abnormal

proteins is different in the spinal cord compared to the CNS regions which are not affected in

ALS. This altered oxidative response may reflect a lower capacity of the spinal motor neurons

to clear/degrade abnormal proteins. This hypothesis is also supported by our identification of

a partially degraded and oxidized form of human mutant SOD1. Oxidation of SOD1 in

Page 80: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

80

general does not result in the toxicity that leads to motor neuron death. This is due to the fact

that mutant SOD1 and human wt SOD1 with same levels of expression as the mutant are

oxidized to the same extent, and moreover, human wt SOD1 over-expression does not

provoke motor neuron disease. However, our data does not completely exclude the possibility

that oxidation renders the mutant SOD1 toxic, because the resulting forms of oxidized wt and

mutant SOD1 may be functionally different. Proteome analysis revealed the upregulation of

PDI, paralleled with increased oxidation of PDI, early in the disease progression that

coincides with accumulation of inactive SOD1 having reduced intrasubunit disulphide bonds.

These results suggest that formation of disulphide bonds and possible thiol-related oxidation

may be altered in familial ALS and may have implications to the SOD1 stability.

SOD1 was destabilized specifically in the spinal cord, which is the most affected

region in ALS pathogenesis. The destabilization, shown as loss of SOD1-dimer stability to

mild denaturaton, was evident throughout the period from late presymptomatic stage to the

end stage. As the redox state of SOD1 may determine the dismutase activity of mutant SOD1,

the nature of the destabilization was further on analyzed by malPEG assay, where malPEG

binds to free cysteine residues and reduced disulphide bonds can be detected. As a positive

control, analysis of the purified human SOD1, when reduced overnight by incubating with

2mM DTT, showed that the reduced, inactive SOD1 had a similar malPEG binding pattern as

the malPEG binding of the mutant SOD1 from the end stage spinal cord. While no changes in

the oxidation state of SOD1 cysteine residues were observed in the SOD1 derived from the

cortices, malPEG reactivity of SOD1 in the spinal cord increased over disease progression,

reaching a statistically significant difference at the end stage of the disease when compared to

SOD1 malPEG reactivity in cortices of the same age. This supports the notion that SOD1

dismutase activity and potential for increased ROS production in the intermembrane space

may be gradually lost, possibly making mutant SOD1 more hydrophobic and more prone to

aggregate, thus shifting the toxic insult from presymptomatic oxidative damage to aggregation

at the symptomatic stages of the disease. Human SOD1 has four cysteine residues and Ferri

and coworkers showed that removal of SOD1 cysteine residues C6 and C111 may suppress

oligomerization of mutant SOD1 by reducing SOD1's ability to form intermolecular

disulphide bonds (Ferri et al., 2006). This is in line with our results showing destabilization

and increased levels of reactive cysteine groups at the symptomatic stages of the disease

progression.

As levels of cysteine reduced SOD1 had increased, we analyzed the levels of protein

disulphide isomerase (PDI) also with Western blotting. PDI is a molecular chaperone, which

functions to rearrange inter- and intramolecular disulphide bonds between cysteine residues of

Page 81: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

81

unfolded proteins (Turano et al., 2002). Human SOD1 has an intramolecular disulphide bond

between cysteine residues C57 and C146 that is required for SOD1 to be active, and thus PDI

may be a key enzyme in controlling SOD1 activity. PDI may activate SOD1 through redox

state of the disulphide bonds, similarly as thioredoxin reductase (Inarrea et al., 2007; Inarrea

et al., 2005). In addition, PDI has been shown to localize to the mitochondria (Rigobello et al.,

2001). PDI was upregulated early in the disease progression and the expression was sharply

decreased towards the end stage, correlating conversely with increased levels of cysteine

reduced SOD1 as shown by malPEG reactivity. PDI upregulation in ALS models has also

been shown by others (Atkin et al., 2006).

Oxidation and destabilization of the mutant SOD1 and reduction of disulphide bonds

within SOD1 combined with upregulation of PDI, a disulphide bond rearranging chaperone,

were specific for the affected spinal cord tissue in rat and mouse models of ALS, indicating a

structural modification that may have an impact on cellular localization, activation and

aggregation of the protein.

6.3 Destabilized mutant SOD1 associates with inner membrane of mitochondria and

increases ROS production (II)

Mutant SOD1 has been shown to be imported preferentially into mitochondria in affected

tissues of SOD1 mouse models (Bergemalm et al., 2006; Deng et al., 2006; Liu et al., 2004;

Vijayvergiya et al., 2005). However, in order for SOD1 to be imported into mitochondria it

must be in a demetallinated state lacking both Cu2+ and Zn2+, and it has to have its

intramolecular disulphide bond reduced (Field et al., 2003). After being imported in to the

intermembrane space of mitochondria, SOD1 is loaded with zinc through an unknown

mechanism and copper is loaded by Cu chaperone for SOD1 (CCS) (Culotta et al., 1997).

After that, SOD1 is kept inactive until activation upon oxidation of its cysteine residues

forming intramolecular disulphide bond (Inarrea et al., 2005).

In order to investigate how destabilized SOD1 may cause mitochondrial dysfunction,

we analyzed mutant SOD1 association to mitochondria. Mutant SOD1 associated to spinal

cord mitochondria to a higher extent than to mitochondria in the cortex of G93A-SOD1 rats.

However, as the mutant SOD1 association to mitochondria may also be related to the over-

expression of the transgene (Bergemalm et al., 2006), we analyzed mutant SOD1 toxicity not

in mitochondria from transgenic rats, but from wt liver mitochondria by exposing mitoplasts

(mitochondria devoid of outer membrane) with cytosolic tissue homogenates of G93A-SOD1

Page 82: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

82

rats from the spinal cord and cortex, with or without destabilized SOD1, respectively. This

model allowed us to re-create the situation where mutant SOD1 is acting in the

intermembrane space and we also were able to distinguish mutant SOD1 from endogenous rat

SOD1 residing in the mitochondrial fraction, as mitochondria and mitoplasts were isolated

from wt rats. The association of destabilized SOD1 from the spinal cord cytosol to wt liver

mitoplasts after exposure was increased and also ROS production was elevated by 28% at the

pre-symptomatic timepoint. The result was reproducible and in our opinion truly noteworthy,

as exposed mitoplasts were from the liver of healthy young wild type rats. Taken together,

exposing mitoplasts to destabilized SOD1 may damage mitochondria, as destabilized mutant

SOD1 bound to the inner membrane of mitochondria and increased ROS production.

Our results indicate that the destabilization with loss of quaternary structure involving

disulphide bonds of mutant SOD1 in the spinal cord makes SOD1 toxic to mitochondria, as

destabilized SOD1 associated to the inner mitochondrial membrane and increased ROS

production. The destabilization may help possibly the mutant SOD1 to enter mitochondria, as

the import of SOD1 requires the reduction of disulphide bonds and demetallination, and the

preferential import of mutant SOD1 may cause toxicity to mitochondria in the intermembrane

space. Still, the mechanism for this toxic function of mutant SOD1 in mitochondria has not

been identified - until now.

6.4 Elevated SOD1 activity in the intermembrane space leads to increased hydrogen

peroxide production resulting in cytochome c catalyzed oxidation (III)

Even though several theories on mechanisms explaining how mutant SOD1 may cause

mitochondrial dysfunction are proposed, including oxidative damage, excitotoxicity with

calcium buffering, and aggregation, there is a lack of conclusive proof of mutant SOD1

toxicity in mitochondria. In fact there is no conclusive proof for one general mechanism of

mutant SOD1 toxicity in ALS at all. What is generally accepted is that many harmful events

like protein aggregation, oxidative damage and excitotoxicity take place in ALS pathogenesis,

occur in concert, and together are combined to ALS disease. What comes to oxidative damage

and mutant SOD1, after the discovery of SOD1 mutations in 1993, mutant SOD1 was

believed to lack activity and the lack of dismutase activity with increased superoxide levels

was thought to be a key contributor for the disease mechanism - this was not the case, as

SOD1 knock out mice do not develop motor neuron disease (Reaume et al., 1996). Secondly,

mutant SOD1 was believed to produce ROSs from aberrant substrates (Beckman et al., 1994;

Page 83: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

83

Estevez et al., 1999; Wiedau-Pazos et al., 1996) - which is most likely not the case as,

increasing the levels of endogenous wild type SOD1 activity did not have any improvement

on the disease pathogenesis in transgenic mouse models. Still, SOD1 activity must play a role,

as increasing the levels of human wt SOD1 expression in mutant SOD1 mice accelerates

disease progression in ALS mice (Deng et al., 2006). From this we come back to the SOD1

activity, not aberrant, but the known accepted SOD1 dismutation activity where superoxide is

converted to oxygen and hydrogen peroxide.

In the cytosol, SOD1 converts superoxide to oxygen and hydrogen peroxide and

hydrogen peroxide is further on converted to water by glutathione peroxidase or catalase as

hydrogen peroxide itself is a strong oxidant. So, in one way of saying, SOD1 takes care of the

job of detoxifying superoxide only halfway, as hydrogen peroxide needs to be cleared by

other enzymes. This is still fine as long as glutathione peroxidase or catalase is present.

However, glutathione peroxidase or catalase are only present in low levels in the

intermembrane space (Martin et al., 1998) and superoxide detoxification is thought to depend

on cytochrome c, which can efficiently oxidize superoxide to oxygen, acting as true

antioxidant (Pereverzev et al., 2003) as none of the products needs to be processed any

further.

Previous studies have demonstrated that the reaction of cytochrome c with hydrogen

peroxide results in the formation of oxoferryl cytochrome c (peroxidase compound I-type

intermediate), which is highly reactive and has a potential to oxidize proteins, DNA and

lipids, as well as endogenous antioxidants such as glutathione, NADH and ascorbate

(Lawrence et al., 2003). In particular, oxidation of cardiolipin, a phospholipid which is in

complex with cytochrome c on the surface of the inner mitochondrial membrane, leads to the

release of proapoptotic factors from mitochondria (Belikova et al., 2006; Kagan et al., 2005).

This leads to a scenario where upon mitochondrial stress, SOD1 might compete with

cytochrome c for superoxide in the intermembrane space and generate hydrogen peroxide,

which then could react with cytochrome c and oxidize the cytochrome c molecule to oxoferryl

heme, a highly reactive oxidant that is able to react with a number of intracellular targets

including proteins, nucleic acids and lipids (Lawrence et al., 2003), eventually leading to a

paradoxical increase in ROS production and cellular injury. The hydrogen peroxide produced

by increased SOD1 activity in the intermembrane space would thus also serve as a substrate

for cardiolipin-bound cytochrome c, and consequently switch on a very early proapoptotic

processes, leading to consecutive programmed cell death.

Our results indicated that upon inhibition of mitochondrial respiration, the elevated

SOD1 activity is responsible for the increased hydrogen peroxide production in the

Page 84: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

84

intermembrane space, resulting in cytochrome c-catalysed oxidation not seen in the

mitochondria isolated from SOD-/- mice. This could trigger a vicious circle where oxidative

damage to mitochondrial respiratory components leads to further ROS production and

peroxidation. Indeed, inhibition of mitochondrial respiration at the level of complex III causes

SOD1-dependent ROS production and apoptotic death of isolated blood lymphocytes.

Moreover, accumulation of mutant human G93A-SOD1 in the intermembrane space that is

observed in tg animal models of ALS, leads to elevated SOD1 activity and increased

cytochrome c-catalyzed oxidation in the intermembrane space.

SOD is generally thought to protect cells from oxidative damage. Accordingly, as a

cytosolic antioxidant SOD1 provides protection in models of transient myocardial (Chen et

al., 2000), brain ischemia (Chan, 2005) and Parkinson's disease (Barkats et al., 2002). Some

other studies, however, suggest that the increased SOD1 activity promotes injury. For

instance, immature mouse brain overexpressing SOD1 shows an increased propensity for

injury and accumulates more hydrogen peroxide after hypoxia-ischemia than wt mouse brain

(Fullerton et al., 1998). Also, elevation of SOD1 increases acoustic trauma from noise

exposure (Endo et al., 2005), and mice deficient in SOD1 are resistant to acetaminophen

toxicity (Lei et al., 2006). Moreover, a superoxide generator, menadione, produces

significantly increased DCF fluorescence and greater death in neurons with mutant SOD1

than in wt neurons, suggesting increased hydrogen peroxide formation in the mutant SOD1

expressing cells (Ying et al., 2000). This apparent discrepancy concerning the role of SOD1 in

cellular injury is explained by the results showing that increased SOD1 activity in the

intermembrane space paradoxically produces peroxides that are converted to highly toxic

ROS.

Mitochondrial dysfunction, including altered function of respiratory complexes, has

been described in arteriosclerosis, diabetes mellitus, and a number of acute and degenerative

brain diseases such as stroke, Parkinson's disease and ALS. Increased ROS production is also

a characteristic of these diseases (Barnham et al., 2004). The increased SOD1 activity

accompanied with high hydrogen peroxide production in the intermembrane space may be

one mechanism of neurodegeneration. Importantly, the toxicity of ALS-linked SOD1 mutants

originates from their selective recruitment to the spinal cord mitochondria (Bergemalm et al.,

2006; Deng et al., 2006; Liu et al., 2004; Vijayvergiya et al., 2005). SOD1 activity in the

intermembrane space may be a relevant therapeutic target for ALS and other degenerative

diseases involving mitochondrial pathogenesis.

Page 85: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

85

6.5 Hypothesized role of destabilized mutant SOD1 toxicity in mitochondria

The reports in the present thesis work provide important information regarding the

possible toxic role of mutant SOD1 in mitochondria in ALS pathogenesis (Figure 5), where

destabilized but active mutant SOD1 is preferentially imported to mitochondria in the spinal

cord. PDI upregulation and conversely correlating levels of reduced cysteine residues of

SOD1 implicate SOD1 inter- and intramolecular disulphide bridges to have a key role in

stability and activity of SOD1. Moreover, destabilized SOD1 may well have implications also

on aggregation, which is a dominant feature of ALS pathogenesis. One striking feature of the

immunoproteasome inhibition by PDTC is that it only occurs in glial cells, namely microglia

and astrocytes, but the damaging effect is seen as reduced survival in rat a model of ALS

despite upregulated GLT-1 levels and anti-inflammatory/oxidant properties of the drug used,

indicating the importance of glial immunoproteasome in ALS pathogenesis. As the glial

immunoproteasome machinery was subdued, the toxicity was focused against motor neurons.

Destabilization of SOD1 may also make it more preferable for mitochondrial import,

as SOD1 needs to be demetallinated and intaramolecular disulphide bond reduced in order to

enter mitochondria. In the mitochondrial intermembrane space, destabilization may also have

a role in SOD1 activation via PDI and production of hydrogen peroxide by activated mutant

SOD1 in the early stages of the disease. In addition to immunoproteasome inhibition, PDTC

also increased intracellular copper concentrations in the CNS, which may lead to higher

SOD1 activity as copper is required in the dismutation reaction, therefore possibly resulting in

increased hydrogen peroxide production. Hydrogen peroxide can react with cytochrome c,

producing highly reactive oxoferryl heme, resulting in oxidation far more damaging than that

of superoxide or hydrogen peroxide, initiating a process of cellular death.

The mechanism presented here for the toxicity of SOD1 in mitochondria provides new

and important information about SOD1 toxicity and enables us to understand better the

underlying factors of motor neuron cell death in ALS, where, SOD1 activation upon stress in

mitochondria and production hydrogen peroxide are the key initiators of oxidative damage

and the process of cellular damage and death. In the future, controlling SOD1 activity in the

intermembrane space or targeting anti-oxidants to the intermembrane space of mitochondria

may be relevant therapeutic approaches for many degenerative diseases involving

mitochondrial pathogenesis, especially ALS.

Page 86: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

86

Figure 5. Hypothesized role of mutant SOD1 in mitochondria and in ALS pathogenesis.Misfolded but active mutant SOD1 localizes to mitochondria, where hydrogen peroxideproduced by SOD1 dismutase activity leads to the formation of oxoferryl heme of cytochromec and oxidation of biological targets. Misfolded SOD1 may well have implications onaggregation as PDI is upregulated in ALS pathogenesis and inhibition of immunoproteasomewith PDTC had adverse effects. PDTC increased copper concentrations and therefore possiblyalso increased SOD1 activity and ROS production. Activation of caspases is the final eventleading to programmed cell death.

SOD1•O2

+ H2O2

cyt coxoferryl-cyt c

H2OO2

cyt c

cyt c

ActiveSOD1

Mitochondria

Proteasome Chaperones

Caspase

Aggregatescontaining SOD1

Immuno-proteasome

PDTC

SOD2•O2–

MisfoldedSOD1

PDI

Cu2+

Page 87: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

87

7. SUMMARY AND CONCLUSIONS

The present study was carried out to study the stability and oxidation of mutant SOD1 and the

toxic role of mutant SOD1 in mitochondria in ALS pathogenesis. In addition, we tested an

anti-inflammatory/oxidant drug with PDTC treatment in an ALS rat model.

PDTC inhibited immunoproteasome induction in glial cells and resulted in decreased

survival despite upregulated GLT-1 and other possible beneficial effects including anti-

inflammatory and anti-oxidative mechanisms by the drug. These results indicate that

immunoproteasome induction in the disease pathogenesis may be an important mechanism for

coping with the toxic consequences of mutant SOD1 in tissues affected in ALS, and also in

the surrounding glial cells. PDTC also increased intracellular copper concentration.

Oxidation of SOD1, destabilization of the quaternary structure with reduction of the

cysteine residues of mutant SOD1 combined with upregulation of PDI, a disulphide bond

rearranging chaperone, were specific for the affected spinal cord tissue in rat and mouse

models of ALS, indicating a structural modification that may have an impact on cellular

localization, activation and aggregation of the protein.

Although SOD1 is classically considered to be a cytosolic enzyme, mutant SOD1

associates to mitochondria in affected spinal cord tissue to a higher extent than to

mitochondria in other regions of the CNS. In addition, destabilized SOD1 associated to the

inner mitochondrial membrane and increased ROS production. The destabilization may

possibly help the mutant SOD1 to enter mitochondria as the import of SOD1 requires

reduction of the disulphide bond and demetallination. Because of the destabilization, mutant

SOD1 is imported preferentially to mitochondria, and may cause damage in the mitochondrial

intermembrane space by interacting with the machinery of the oxidative phosphorylation.

The mechanism of this interaction was shown to be a situation where upon

mitochondrial stress SOD1 might compete with cytochrome c for superoxide in the

mitochondrial intermembrane space and generate hydrogen peroxide, which then could react

with cytochrome c and oxidize the cytochrome c molecule to oxoferryl heme. Cytochrome c

oxidized to oxoferryl heme is a highly reactive oxidant, far more reactive than superoxide or

hydrogen peroxide, capable of rapidly reacting with a number of intracellular targets

including proteins, nucleic acids and lipids, eventually leading to a paradoxical increase in

ROS production and cellular injury leading to consecutive programmed cell death.

The mechanism presented here for the toxicity of SOD1 in mitochondria provides new

and important information about SOD1 toxicity and enables us to understand better the

underlying factors of motor neuron cell death in ALS, where SOD1 activation upon stress in

Page 88: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

88

mitochondria and production of hydrogen peroxide are the key initiators of oxidative damage

and the process of cellular damage and death. In the future, controlling SOD1 activity in the

intermembrane space or targeting anti-oxidants to the intermembrane space of mitochondria

may be relevant therapeutic approaches for many degenerative diseases involving

mitochondrial pathogenesis, especially ALS.

Page 89: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

89

8. REFERENCES

Afifi, A. K., Aleu, F. P., Goodgold, J., and MacKay, B. (1966). Ultrastructure of atrophicmuscle in amyotrophic lateral sclerosis. Neurology 16, 475-481.

Al-Chalabi, A., Andersen, P. M., Chioza, B., Shaw, C., Sham, P. C., Robberecht, W.,Matthijs, G., Camu, W., Marklund, S. L., Forsgren, L., et al. (1998). Recessiveamyotrophic lateral sclerosis families with the D90A SOD1 mutation share a commonfounder: evidence for a linked protective factor. Hum Mol Genet 7, 2045-2050.

Alexianu, M. E., Kozovska, M., and Appel, S. H. (2001). Immune reactivity in a mousemodel of familial ALS correlates with disease progression. Neurology 57, 1282-1289.

Almer, G., Teismann, P., Stevic, Z., Halaschek-Wiener, J., Deecke, L., Kostic, V., andPrzedborski, S. (2002). Increased levels of the pro-inflammatory prostaglandin PGE2in CSF from ALS patients. Neurology 58, 1277-1279.

ALS CNTF Treatment Study Group (1996). A double-blind placebo-controlled clinical trialof subcutaneous recombinant human ciliary neurotrophic factor (rHCNTF) inamyotrophic lateral sclerosis. ALS CNTF Treatment Study Group. Neurology 46,1244-1249.

Alzheimer, A. (1907a). Uber eine eigenartige Erkankung der Hirnrinde. Psychiatr Psych-Gerichtl Med 64, 146-148.

Alzheimer, A. (1907b). Uber eine eigenartige Erkankung der Hirnrinde. ZentralblNervenheilkd Psychiatr 30, 177-179.

Andersen, P. M., Forsgren, L., Binzer, M., Nilsson, P., Ala-Hurula, V., Keranen, M. L.,Bergmark, L., Saarinen, A., Haltia, T., Tarvainen, I., et al. (1996). Autosomalrecessive adult-onset amyotrophic lateral sclerosis associated with homozygosity forAsp90Ala CuZn-superoxide dismutase mutation. A clinical and genealogical study of36 patients. Brain 119 ( Pt 4), 1153-1172.

Andersen, P. M., Nilsson, P., Ala-Hurula, V., Keranen, M. L., Tarvainen, I., Haltia, T.,Nilsson, L., Binzer, M., Forsgren, L., and Marklund, S. L. (1995). Amyotrophic lateralsclerosis associated with homozygosity for an Asp90Ala mutation in CuZn-superoxidedismutase. Nat Genet 10, 61-66.

Andrews, J. M., Gardner, M. B., Wolfgram, F. J., Ellison, G. W., Porter, D. D., andBrandkamp, W. W. (1974). Studies on a murine form of spontaneous lower motorneuron degeneration--the wobbler (wa) mouse. Am J Pathol 76, 63-78.

Andrus, P. K., Fleck, T. J., Gurney, M. E., and Hall, E. D. (1998). Protein oxidative damagein a transgenic mouse model of familial amyotrophic lateral sclerosis. J Neurochem71, 2041-2048.

Annis, C. M., and Vaughn, J. E. (1998). Differential vulnerability of autonomic and somaticmotor neurons to N-methyl-D-aspartate-induced excitotoxicity. Neuroscience 83, 239-249.

Aoki, M., Ogasawara, M., Matsubara, Y., Narisawa, K., Nakamura, S., Itoyama, Y., and Abe,K. (1993). Mild ALS in Japan associated with novel SOD mutation. Nat Genet 5, 323-324.

Atkin, J. D., Farg, M. A., Turner, B. J., Tomas, D., Lysaght, J. A., Nunan, J., Rembach, A.,Nagley, P., Beart, P. M., Cheema, S. S., and Horne, M. K. (2006). Induction of theunfolded protein response in familial amyotrophic lateral sclerosis and association ofprotein-disulfide isomerase with superoxide dismutase 1. J Biol Chem 281, 30152-30165.

Azzouz, M., Ralph, G. S., Storkebaum, E., Walmsley, L. E., Mitrophanous, K. A., Kingsman,S. M., Carmeliet, P., and Mazarakis, N. D. (2004). VEGF delivery with retrogradelytransported lentivector prolongs survival in a mouse ALS model. Nature 429, 413-417.

Page 90: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

90

Barkats, M., Millecamps, S., Bilang-Bleuel, A., and Mallet, J. (2002). Neuronal transfer of thehuman Cu/Zn superoxide dismutase gene increases the resistance of dopaminergicneurons to 6-hydroxydopamine. J Neurochem 82, 101-109.

Barnham, K. J., Masters, C. L., and Bush, A. I. (2004). Neurodegenerative diseases andoxidative stress. Nat Rev Drug Discov 3, 205-214.

Batulan, Z., Shinder, G. A., Minotti, S., He, B. P., Doroudchi, M. M., Nalbantoglu, J., Strong,M. J., and Durham, H. D. (2003). High threshold for induction of the stress responsein motor neurons is associated with failure to activate HSF1. J Neurosci 23, 5789-5798.

Baumeister, W., Walz, J., Zuhl, F., and Seemuller, E. (1998). The proteasome: paradigm of aself-compartmentalizing protease. Cell 92, 367-380.

BDNF Study Group (1999). A controlled trial of recombinant methionyl human BDNF inALS: The BDNF Study Group (Phase III). Neurology 52, 1427-1433.

Beal, M. F., Ferrante, R. J., Browne, S. E., Matthews, R. T., Kowall, N. W., and Brown, R.H., Jr. (1997). Increased 3-nitrotyrosine in both sporadic and familial amyotrophiclateral sclerosis. Ann Neurol 42, 644-654.

Beauchamp, C., and Fridovich, I. (1971). Superoxide dismutase: improved assays and anassay applicable to acrylamide gels. Anal Biochem 44, 276-287.

Beckman, J. S., Carson, M., Smith, C. D., and Koppenol, W. H. (1993). ALS, SOD andperoxynitrite. Nature 364, 584.

Beckman, J. S., Chen, J., Crow, J. P., and Ye, Y. Z. (1994). Reactions of nitric oxide,superoxide and peroxynitrite with superoxide dismutase in neurodegeneration. ProgBrain Res 103, 371-380.

Beers, D. R., Henkel, J. S., Xiao, Q., Zhao, W., Wang, J., Yen, A. A., Siklos, L., McKercher,S. R., and Appel, S. H. (2006). Wild-type microglia extend survival in PU.1 knockoutmice with familial amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A 103,16021-16026.

Belikova, N. A., Vladimirov, Y. A., Osipov, A. N., Kapralov, A. A., Tyurin, V. A.,Potapovich, M. V., Basova, L. V., Peterson, J., Kurnikov, I. V., and Kagan, V. E.(2006). Peroxidase Activity and Structural Transitions of Cytochrome c Bound toCardiolipin-Containing Membranes. Biochemistry 45, 4998-5009.

Bensimon, G., Lacomblez, L., and Meininger, V. (1994). A controlled trial of riluzole inamyotrophic lateral sclerosis. ALS/Riluzole Study Group. N Engl J Med 330, 585-591.

Bergemalm, D., Jonsson, P. A., Graffmo, K. S., Andersen, P. M., Brannstrom, T., Rehnmark,A., and Marklund, S. L. (2006). Overloading of stable and exclusion of unstablehuman superoxide dismutase-1 variants in mitochondria of murine amyotrophic lateralsclerosis models. J Neurosci 26, 4147-4154.

Blauw, H. M., Veldink, J. H., van Es, M. A., van Vught, P. W., Saris, C. G., van der Zwaag,B., Franke, L., Burbach, J. P., Wokke, J. H., Ophoff, R. A., and van den Berg, L. H.(2008). Copy-number variation in sporadic amyotrophic lateral sclerosis: a genome-wide screen. Lancet Neurol 7, 319-326.

Boillee, S., Vande Velde, C., and Cleveland, D. W. (2006a). ALS: A Disease of MotorNeurons and Their Nonneuronal Neighbors. Neuron 52, 39-59.

Boillee, S., Yamanaka, K., Lobsiger, C. S., Copeland, N. G., Jenkins, N. A., Kassiotis, G.,Kollias, G., and Cleveland, D. W. (2006b). Onset and progression in inherited ALSdetermined by motor neurons and microglia. Science 312, 1389-1392.

Bommel, H., Xie, G., Rossoll, W., Wiese, S., Jablonka, S., Boehm, T., and Sendtner, M.(2002). Missense mutation in the tubulin-specific chaperone E (Tbce) gene in themouse mutant progressive motor neuronopathy, a model of human motoneurondisease. J Cell Biol 159, 563-569.

Page 91: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

91

Borasio, G. D., Robberecht, W., Leigh, P. N., Emile, J., Guiloff, R. J., Jerusalem, F., Silani,V., Vos, P. E., Wokke, J. H., and Dobbins, T. (1998). A placebo-controlled trial ofinsulin-like growth factor-I in amyotrophic lateral sclerosis. European ALS/IGF-IStudy Group. Neurology 51, 583-586.

Borchelt, D. R., Lee, M. K., Slunt, H. S., Guarnieri, M., Xu, Z. S., Wong, P. C., Brown, R. H.,Jr., Price, D. L., Sisodia, S. S., and Cleveland, D. W. (1994). Superoxide dismutase 1with mutations linked to familial amyotrophic lateral sclerosis possesses significantactivity. Proc Natl Acad Sci U S A 91, 8292-8296.

Boston-Howes, W., Gibb, S. L., Williams, E. O., Pasinelli, P., Brown, R. H., Jr., and Trotti,D. (2006). Caspase-3 cleaves and inactivates the glutamate transporter EAAT2. J BiolChem 281, 14076-14084.

Bowling, A. C., Barkowski, E. E., McKenna-Yasek, D., Sapp, P., Horvitz, H. R., Beal, M. F.,and Brown, R. H., Jr. (1995). Superoxide dismutase concentration and activity infamilial amyotrophic lateral sclerosis. J Neurochem 64, 2366-2369.

Bowling, A. C., Schulz, J. B., Brown, R. H., Jr., and Beal, M. F. (1993). Superoxidedismutase activity, oxidative damage, and mitochondrial energy metabolism infamilial and sporadic amyotrophic lateral sclerosis. J Neurochem 61, 2322-2325.

Brockington, A., Kirby, J., Eggitt, D., Schofield, E., Morris, C., Lewis, C. E., Ince, P. G., andShaw, P. J. (2005). Screening of the regulatory and coding regions of vascularendothelial growth factor in amyotrophic lateral sclerosis. Neurogenetics 6, 101-104.

Broom, W. J., Russ, C., Sapp, P. C., McKenna-Yasek, D., Hosler, B. A., Andersen, P. M., andBrown, R. H., Jr. (2006). Variants in candidate ALS modifier genes linked to Cu/Znsuperoxide dismutase do not explain divergent survival phenotypes. Neurosci Lett392, 52-57.

Browne, S. E., Yang, L., Dimauro, J. P., Fuller, S. W., Licata, S. C., and Beal, M. F. (2006).Bioenergetic abnormalities in discrete cerebral motor pathways presage spinal cordpathology in the G93A SOD1 mouse model of ALS. Neurobiol Dis.

Bruening, W., Roy, J., Giasson, B., Figlewicz, D. A., Mushynski, W. E., and Durham, H. D.(1999). Up-regulation of protein chaperones preserves viability of cells expressingtoxic Cu/Zn-superoxide dismutase mutants associated with amyotrophic lateralsclerosis. J Neurochem 72, 693-699.

Bruijn, L. I., Beal, M. F., Becher, M. W., Schulz, J. B., Wong, P. C., Price, D. L., andCleveland, D. W. (1997a). Elevated free nitrotyrosine levels, but not protein-boundnitrotyrosine or hydroxyl radicals, throughout amyotrophic lateral sclerosis (ALS)-likedisease implicate tyrosine nitration as an aberrant in vivo property of one familialALS-linked superoxide dismutase 1 mutant. Proc Natl Acad Sci U S A 94, 7606-7611.

Bruijn, L. I., Becher, M. W., Lee, M. K., Anderson, K. L., Jenkins, N. A., Copeland, N. G.,Sisodia, S. S., Rothstein, J. D., Borchelt, D. R., Price, D. L., and Cleveland, D. W.(1997b). ALS-linked SOD1 mutant G85R mediates damage to astrocytes andpromotes rapidly progressive disease with SOD1-containing inclusions. Neuron 18,327-338.

Bruijn, L. I., Houseweart, M. K., Kato, S., Anderson, K. L., Anderson, S. D., Ohama, E.,Reaume, A. G., Scott, R. W., and Cleveland, D. W. (1998). Aggregation and motorneuron toxicity of an ALS-linked SOD1 mutant independent from wild-type SOD1.Science 281, 1851-1854.

Bruijn, L. I., Miller, T. M., and Cleveland, D. W. (2004a). Supplemental Material: Unravelingthe Mechanisms Involved in Motor Neuron Degeneration in ALS. Evaluation ofputative therapeutics in mice and humans. Annu Rev Neurosci 27, 723-749.

Bruijn, L. I., Miller, T. M., and Cleveland, D. W. (2004b). Unraveling the MechanismsInvolved in Motor Neuron Degeneration in Als. Annu Rev Neurosci 27, 723-749.

Cai, H., Lin, X., Xie, C., Laird, F. M., Lai, C., Wen, H., Chiang, H. C., Shim, H., Farah, M.H., Hoke, A., et al. (2005). Loss of ALS2 function is insufficient to trigger motor

Page 92: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

92

neuron degeneration in knock-out mice but predisposes neurons to oxidative stress. JNeurosci 25, 7567-7574.

Calviello, G., Filippi, G. M., Toesca, A., Palozza, P., Maggiano, N., Nicuolo, F. D., Serini, S.,Azzena, G. B., and Galeotti, T. (2005). Repeated exposure to pyrrolidine-dithiocarbamate induces peripheral nerve alterations in rats. Toxicol Lett 158, 61-71.

Camu, W., Khoris, J., Moulard, B., Salachas, F., Briolotti, V., Rouleau, G. A., and Meininger,V. (1999). Genetics of familial ALS and consequences for diagnosis. French ALSResearch Group. J Neurol Sci 165 Suppl 1, S21-26.

Chan, P. H. (2005). Mitochondrial dysfunction and oxidative stress as determinants of celldeath/survival in stroke. Ann N Y Acad Sci 1042, 203-209.

Chan, P. H., Kawase, M., Murakami, K., Chen, S. F., Li, Y., Calagui, B., Reola, L., Carlson,E., and Epstein, C. J. (1998). Overexpression of SOD1 in transgenic rats protectsvulnerable neurons against ischemic damage after global cerebral ischemia andreperfusion. J Neurosci 18, 8292-8299.

Chance, P. F., Rabin, B. A., Ryan, S. G., Ding, Y., Scavina, M., Crain, B., Griffin, J. W., andCornblath, D. R. (1998). Linkage of the gene for an autosomal dominant form ofjuvenile amyotrophic lateral sclerosis to chromosome 9q34. Am J Hum Genet 62, 633-640.

Charcot, J. M., and Joffory, A. (1869). Deux cas d'atrophie musculaire progressive aveclesions de la substance grise et des faisceaux antero-lateraux de la moelle epiniere.Arch Physiol Neurol Pathol 2, 744-754.

Chen, K., Long, Y. M., Wang, H., Lan, L., and Lin, Z. H. (2005). Activation of nuclearfactor-kappa B and effects of pyrrolidine dithiocarbamate on TNBS-induced ratcolitis. World J Gastroenterol 11, 1508-1514.

Chen, Y. Z., Bennett, C. L., Huynh, H. M., Blair, I. P., Puls, I., Irobi, J., Dierick, I., Abel, A.,Kennerson, M. L., Rabin, B. A., et al. (2004). DNA/RNA helicase gene mutations in aform of juvenile amyotrophic lateral sclerosis (ALS4). Am J Hum Genet 74, 1128-1135.

Chen, Z., Oberley, T. D., Ho, Y., Chua, C. C., Siu, B., Hamdy, R. C., Epstein, C. J., andChua, B. H. (2000). Overexpression of CuZnSOD in coronary vascular cellsattenuates myocardial ischemia/reperfusion injury. Free Radic Biol Med 29, 589-596.

Cheng, G., Whitehead, S. N., Hachinski, V., and Cechetto, D. F. (2006). Effects of pyrrolidinedithiocarbamate on beta-amyloid (25-35)-induced inflammatory responses andmemory deficits in the rat. Neurobiol Dis.

Cheroni, C., Peviani, M., Cascio, P., Debiasi, S., Monti, C., and Bendotti, C. (2005).Accumulation of human SOD1 and ubiquitinated deposits in the spinal cord ofSOD1G93A mice during motor neuron disease progression correlates with a decreaseof proteasome. Neurobiol Dis 18, 509-522.

Chi, L., Ke, Y., Luo, C., Li, B., Gozal, D., Kalyanaraman, B., and Liu, R. (2006). Motorneuron degeneration promotes neural progenitor cell proliferation, migration, andneurogenesis in the spinal cords of amyotrophic lateral sclerosis mice. Stem Cells 24,34-43.

Cifuentes-Diaz, C., Frugier, T., Tiziano, F. D., Lacene, E., Roblot, N., Joshi, V., Moreau, M.H., and Melki, J. (2001). Deletion of murine SMN exon 7 directed to skeletal muscleleads to severe muscular dystrophy. J Cell Biol 152, 1107-1114.

Clement, A. M., Nguyen, M. D., Roberts, E. A., Garcia, M. L., Boillee, S., Rule, M.,McMahon, A. P., Doucette, W., Siwek, D., Ferrante, R. J., et al. (2003). Wild-typenonneuronal cells extend survival of SOD1 mutant motor neurons in ALS mice.Science 302, 113-117.

Cleveland, D. W., Laing, N., Hurse, P. V., and Brown, R. H., Jr. (1995). Toxic mutants inCharcot's sclerosis. Nature 378, 342-343.

Page 93: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

93

Cleveland, D. W., and Rothstein, J. D. (2001). From Charcot to Lou Gehrig: decipheringselective motor neuron death in ALS. Nat Rev Neurosci 2, 806-819.

Cook, S. A., Johnson, K. R., Bronson, R. T., and Davisson, M. T. (1995). Neuromusculardegeneration (nmd): a mutation on mouse chromosome 19 that causes motor neurondegeneration. Mamm Genome 6, 187-191.

Corti, S., Locatelli, F., Donadoni, C., Guglieri, M., Papadimitriou, D., Strazzer, S., Del Bo,R., and Comi, G. P. (2004). Wild-type bone marrow cells ameliorate the phenotype ofSOD1-G93A ALS mice and contribute to CNS, heart and skeletal muscle tissues.Brain 127, 2518-2532.

Cox, G. A., Mahaffey, C. L., and Frankel, W. N. (1998). Identification of the mouseneuromuscular degeneration gene and mapping of a second site suppressor allele.Neuron 21, 1327-1337.

Cudkowicz, M. E., McKenna-Yasek, D., Sapp, P. E., Chin, W., Geller, B., Hayden, D. L.,Schoenfeld, D. A., Hosler, B. A., Horvitz, H. R., and Brown, R. H. (1997).Epidemiology of mutations in superoxide dismutase in amyotrophic lateral sclerosis.Ann Neurol 41, 210-221.

Cudkowicz, M. E., Shefner, J. M., Schoenfeld, D. A., Zhang, H., Andreasson, K. I., Rothstein,J. D., and Drachman, D. B. (2006). Trial of celecoxib in amyotrophic lateral sclerosis.Ann Neurol 60, 22-31.

Culotta, V. C., Klomp, L. W., Strain, J., Casareno, R. L., Krems, B., and Gitlin, J. D. (1997).The copper chaperone for superoxide dismutase. J Biol Chem 272, 23469-23472.

Cuzzocrea, S., Chatterjee, P. K., Mazzon, E., Dugo, L., Serraino, I., Britti, D., Mazzullo, G.,Caputi, A. P., and Thiemermann, C. (2002). Pyrrolidine dithiocarbamate attenuates thedevelopment of acute and chronic inflammation. Br J Pharmacol 135, 496-510.

Dal Canto, M. C., and Gurney, M. E. (1994). Development of central nervous systempathology in a murine transgenic model of human amyotrophic lateral sclerosis. Am JPathol 145, 1271-1279.

Damiano, M., Starkov, A. A., Petri, S., Kipiani, K., Kiaei, M., Mattiazzi, M., Flint Beal, M.,and Manfredi, G. (2006). Neural mitochondrial Ca2+ capacity impairment precedesthe onset of motor symptoms in G93A Cu/Zn-superoxide dismutase mutant mice. JNeurochem 96, 1349-1361.

Delfs, J. R., Saroff, D. M., Nishida, Y., Friend, J., and Geula, C. (1997). Effects of NMDAand its antagonists on ventral horn cholinergic neurons in organotypic roller tubespinal cord cultures. J Neural Transm 104, 31-51.

DeMartino, G. N., and Slaughter, C. A. (1999). The proteasome, a novel protease regulated bymultiple mechanisms. J Biol Chem 274, 22123-22126.

Deng, H. X., Hentati, A., Tainer, J. A., Iqbal, Z., Cayabyab, A., Hung, W. Y., Getzoff, E. D.,Hu, P., Herzfeldt, B., Roos, R. P., and et al. (1993). Amyotrophic lateral sclerosis andstructural defects in Cu,Zn superoxide dismutase. Science 261, 1047-1051.

Deng, H. X., Shi, Y., Furukawa, Y., Zhai, H., Fu, R., Liu, E., Gorrie, G. H., Khan, M. S.,Hung, W. Y., Bigio, E. H., et al. (2006). Conversion to the amyotrophic lateralsclerosis phenotype is associated with intermolecular linked insoluble aggregates ofSOD1 in mitochondria. Proc Natl Acad Sci U S A 103, 7142-7147.

Deshpande, D. M., Kim, Y. S., Martinez, T., Carmen, J., Dike, S., Shats, I., Rubin, L. L.,Drummond, J., Krishnan, C., Hoke, A., et al. (2006). Recovery from paralysis in adultrats using embryonic stem cells. Ann Neurol 60, 32-44.

Di Giorgio, F. P., Carrasco, M. A., Siao, M. C., Maniatis, T., and Eggan, K. (2007). Non-cellautonomous effect of glia on motor neurons in an embryonic stem cell-based ALSmodel. Nat Neurosci 10, 608-614.

Dick, T. P., Ruppert, T., Groettrup, M., Kloetzel, P. M., Kuehn, L., Koszinowski, U. H.,Stevanovic, S., Schild, H., and Rammensee, H. G. (1996). Coordinated dual cleavages

Page 94: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

94

induced by the proteasome regulator PA28 lead to dominant MHC ligands. Cell 86,253-262.

Donaghy, M. (1999). Classification and clinical features of motor neurone diseases and motorneuropathies in adults. J Neurol 246, 331-333.

Drachman, D. B., Frank, K., Dykes-Hoberg, M., Teismann, P., Almer, G., Przedborski, S.,and Rothstein, J. D. (2002). Cyclooxygenase 2 inhibition protects motor neurons andprolongs survival in a transgenic mouse model of ALS. Ann Neurol 52, 771-778.

Durham, H. D., Roy, J., Dong, L., and Figlewicz, D. A. (1997). Aggregation of mutant Cu/Znsuperoxide dismutase proteins in a culture model of ALS. J Neuropathol Exp Neurol56, 523-530.

Elia, A. J., Parkes, T. L., Kirby, K., St George-Hyslop, P., Boulianne, G. L., Phillips, J. P.,and Hilliker, A. J. (1999). Expression of human FALS SOD in motorneurons ofDrosophila. Free Radic Biol Med 26, 1332-1338.

Elliott, J. L. (1999). Experimental models of amyotrophic lateral sclerosis. Neurobiol Dis 6,310-320.

Elliott, J. L. (2001). Cytokine upregulation in a murine model of familial amyotrophic lateralsclerosis. Brain Res Mol Brain Res 95, 172-178.

Enayat, Z. E., Orrell, R. W., Claus, A., Ludolph, A., Bachus, R., Brockmuller, J., Ray-Chaudhuri, K., Radunovic, A., Shaw, C., Wilkinson, J., and et al. (1995). Two novelmutations in the gene for copper zinc superoxide dismutase in UK families withamyotrophic lateral sclerosis. Hum Mol Genet 4, 1239-1240.

Endo, T., Nakagawa, T., Iguchi, F., Kita, T., Okano, T., Sha, S. H., Schacht, J., Shiga, A.,Kim, T. S., and Ito, J. (2005). Elevation of superoxide dismutase increases acoustictrauma from noise exposure. Free Radic Biol Med 38, 492-498.

Engelhardt, J. I., and Appel, S. H. (1990). IgG reactivity in the spinal cord and motor cortexin amyotrophic lateral sclerosis. Arch Neurol 47, 1210-1216.

Estevez, A. G., Crow, J. P., Sampson, J. B., Reiter, C., Zhuang, Y., Richardson, G. J., Tarpey,M. M., Barbeito, L., and Beckman, J. S. (1999). Induction of nitric oxide-dependentapoptosis in motor neurons by zinc-deficient superoxide dismutase. Science 286,2498-2500.

Ferrante, R. J., Shinobu, L. A., Schulz, J. B., Matthews, R. T., Thomas, C. E., Kowall, N. W.,Gurney, M. E., and Beal, M. F. (1997). Increased 3-nitrotyrosine and oxidativedamage in mice with a human copper/zinc superoxide dismutase mutation. AnnNeurol 42, 326-334.

Ferri, A., Cozzolino, M., Crosio, C., Nencini, M., Casciati, A., Gralla, E. B., Rotilio, G.,Valentine, J. S., and Carri, M. T. (2006). Familial ALS-superoxide dismutasesassociate with mitochondria and shift their redox potentials. Proc Natl Acad Sci U S A103, 13860-13865.

Field, L. S., Furukawa, Y., O'Halloran, T. V., and Culotta, V. C. (2003). Factors controllingthe uptake of yeast copper/zinc superoxide dismutase into mitochondria. J Biol Chem278, 28052-28059.

Fischer, L. R., Culver, D. G., Tennant, P., Davis, A. A., Wang, M., Castellano-Sanchez, A.,Khan, J., Polak, M. A., and Glass, J. D. (2004). Amyotrophic lateral sclerosis is adistal axonopathy: evidence in mice and man. Exp Neurol 185, 232-240.

Frey, D., Schneider, C., Xu, L., Borg, J., Spooren, W., and Caroni, P. (2000). Early andselective loss of neuromuscular synapse subtypes with low sprouting competence inmotoneuron diseases. J Neurosci 20, 2534-2542.

Fridovich, I. (1986). Superoxide dismutase. Adv Enzymol 58, 61-97.Fruh, K., Gossen, M., Wang, K., Bujard, H., Peterson, P. A., and Yang, Y. (1994).

Displacement of housekeeping proteasome subunits by MHC-encoded LMPs: a newlydiscovered mechanism for modulating the multicatalytic proteinase complex. Embo J13, 3236-3244.

Page 95: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

95

Fullerton, H. J., Ditelberg, J. S., Chen, S. F., Sarco, D. P., Chan, P. H., Epstein, C. J., andFerriero, D. M. (1998). Copper/zinc superoxide dismutase transgenic brainaccumulates hydrogen peroxide after perinatal hypoxia ischemia. Ann Neurol 44, 357-364.

Garbuzova-Davis, S., Willing, A. E., Zigova, T., Saporta, S., Justen, E. B., Lane, J. C.,Hudson, J. E., Chen, N., Davis, C. D., and Sanberg, P. R. (2003). Intravenousadministration of human umbilical cord blood cells in a mouse model of amyotrophiclateral sclerosis: distribution, migration, and differentiation. J Hematother Stem CellRes 12, 255-270.

Gong, Y. H., Parsadanian, A. S., Andreeva, A., Snider, W. D., and Elliott, J. L. (2000).Restricted expression of G86R Cu/Zn superoxide dismutase in astrocytes results inastrocytosis but does not cause motoneuron degeneration. J Neurosci 20, 660-665.

Gordon, P. H., Moore, D. H., Miller, R. G., Florence, J. M., Verheijde, J. L., Doorish, C.,Hilton, J. F., Spitalny, G. M., MacArthur, R. B., Mitsumoto, H., et al. (2007). Efficacyof minocycline in patients with amyotrophic lateral sclerosis: a phase III randomisedtrial. Lancet Neurol 6, 1045-1053.

Gould, T. W., Buss, R. R., Vinsant, S., Prevette, D., Sun, W., Knudson, C. M., Milligan, C.E., and Oppenheim, R. W. (2006). Complete dissociation of motor neuron death frommotor dysfunction by Bax deletion in a mouse model of ALS. J Neurosci 26, 8774-8786.

Greenway, M. J., Andersen, P. M., Russ, C., Ennis, S., Cashman, S., Donaghy, C., Patterson,V., Swingler, R., Kieran, D., Prehn, J., et al. (2006). ANG mutations segregate withfamilial and 'sporadic' amyotrophic lateral sclerosis. Nat Genet 38, 411-413.

Grimes, C. A., and Jope, R. S. (2001). The multifaceted roles of glycogen synthase kinase3beta in cellular signaling. Prog Neurobiol 65, 391-426.

Groeneveld, G. J., Veldink, J. H., van der Tweel, I., Kalmijn, S., Beijer, C., de Visser, M.,Wokke, J. H., Franssen, H., and van den Berg, L. H. (2003). A randomized sequentialtrial of creatine in amyotrophic lateral sclerosis. Ann Neurol 53, 437-445.

Gros-Louis, F., Gaspar, C., and Rouleau, G. A. (2006). Genetics of familial and sporadicamyotrophic lateral sclerosis. Biochim Biophys Acta 1762, 956-972.

Gros-Louis, F., Laurent, S., Lopes, A. A., Khoris, J., Meininger, V., Camu, W., and Rouleau,G. A. (2003). Absence of mutations in the hypoxia response element of VEGF inALS. Muscle Nerve 28, 774-775.

Gurney, M. E., Cutting, F. B., Zhai, P., Doble, A., Taylor, C. P., Andrus, P. K., and Hall, E.D. (1996). Benefit of vitamin E, riluzole, and gabapentin in a transgenic model offamilial amyotrophic lateral sclerosis. Ann Neurol 39, 147-157.

Gurney, M. E., Pu, H., Chiu, A. Y., Dal Canto, M. C., Polchow, C. Y., Alexander, D. D.,Caliendo, J., Hentati, A., Kwon, Y. W., Deng, H. X., and et al. (1994). Motor neurondegeneration in mice that express a human Cu,Zn superoxide dismutase mutation.Science 264, 1772-1775.

Hadano, S., Hand, C. K., Osuga, H., Yanagisawa, Y., Otomo, A., Devon, R. S., Miyamoto,N., Showguchi-Miyata, J., Okada, Y., Singaraja, R., et al. (2001). A gene encoding aputative GTPase regulator is mutated in familial amyotrophic lateral sclerosis 2. NatGenet 29, 166-173.

Hall, E. D., Andrus, P. K., Oostveen, J. A., Fleck, T. J., and Gurney, M. E. (1998a).Relationship of oxygen radical-induced lipid peroxidative damage to disease onset andprogression in a transgenic model of familial ALS. J Neurosci Res 53, 66-77.

Hall, E. D., Oostveen, J. A., and Gurney, M. E. (1998b). Relationship of microglial andastrocytic activation to disease onset and progression in a transgenic model of familialALS. Glia 23, 249-256.

Page 96: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

96

Han, D., Antunes, F., Canali, R., Rettori, D., and Cadenas, E. (2003). Voltage-dependentanion channels control the release of the superoxide anion from mitochondria tocytosol. J Biol Chem 278, 5557-5563.

Hand, C. K., Khoris, J., Salachas, F., Gros-Louis, F., Lopes, A. A., Mayeux-Portas, V.,Brewer, C. G., Brown, R. H., Jr., Meininger, V., Camu, W., and Rouleau, G. A.(2002). A novel locus for familial amyotrophic lateral sclerosis, on chromosome 18q.Am J Hum Genet 70, 251-256.

Hanisch, U. K. (2002). Microglia as a source and target of cytokines. Glia 40, 140-155.Hanson, M. G., Jr., Shen, S., Wiemelt, A. P., McMorris, F. A., and Barres, B. A. (1998).

Cyclic AMP elevation is sufficient to promote the survival of spinal motor neurons invitro. J Neurosci 18, 7361-7371.

Harding, A. E., and Thomas, P. K. (1980). Hereditary distal spinal muscular atrophy. A reporton 34 cases and a review of the literature. J Neurol Sci 45, 337-348.

Harper, J. M., Krishnan, C., Darman, J. S., Deshpande, D. M., Peck, S., Shats, I., Backovic,S., Rothstein, J. D., and Kerr, D. A. (2004). Axonal growth of embryonic stem cell-derived motoneurons in vitro and in motoneuron-injured adult rats. Proc Natl Acad SciU S A 101, 7123-7128.

Harraz, M. M., Marden, J. J., Zhou, W., Zhang, Y., Williams, A., Sharov, V. S., Nelson, K.,Luo, M., Paulson, H., Schoneich, C., and Engelhardt, J. F. (2008). SOD1 mutationsdisrupt redox-sensitive Rac regulation of NADPH oxidase in a familial ALS model. JClin Invest 118, 659-670.

Haverkamp, L. J., Appel, V., and Appel, S. H. (1995). Natural history of amyotrophic lateralsclerosis in a database population. Validation of a scoring system and a model forsurvival prediction. Brain 118 ( Pt 3), 707-719.

Hayakawa, M., Miyashita, H., Sakamoto, I., Kitagawa, M., Tanaka, H., Yasuda, H., Karin,M., and Kikugawa, K. (2003). Evidence that reactive oxygen species do not mediateNF-kappaB activation. Embo J 22, 3356-3366.

Helenius, M., Hanninen, M., Lehtinen, S. K., and Salminen, A. (1996). Changes associatedwith aging and replicative senescence in the regulation of transcription factor nuclearfactor-kappa B. Biochem J 318 ( Pt 2), 603-608.

Hempel, S. L., Buettner, G. R., O'Malley, Y. Q., Wessels, D. A., and Flaherty, D. M. (1999).Dihydrofluorescein diacetate is superior for detecting intracellular oxidants:comparison with 2',7'-dichlorodihydrofluorescein diacetate, 5(and 6)-carboxy-2',7'-dichlorodihydrofluorescein diacetate, and dihydrorhodamine 123. Free Radic BiolMed 27, 146-159.

Henkel, J. S., Beers, D. R., Siklos, L., and Appel, S. H. (2006). The chemokine MCP-1 andthe dendritic and myeloid cells it attracts are increased in the mSOD1 mouse model ofALS. Mol Cell Neurosci 31, 427-437.

Henkel, J. S., Engelhardt, J. I., Siklos, L., Simpson, E. P., Kim, S. H., Pan, T., Goodman, J.C., Siddique, T., Beers, D. R., and Appel, S. H. (2004). Presence of dendritic cells,MCP-1, and activated microglia/macrophages in amyotrophic lateral sclerosis spinalcord tissue. Ann Neurol 55, 221-235.

Hensley, K., Floyd, R. A., Gordon, B., Mou, S., Pye, Q. N., Stewart, C., West, M., andWilliamson, K. (2002). Temporal patterns of cytokine and apoptosis-related geneexpression in spinal cords of the G93A-SOD1 mouse model of amyotrophic lateralsclerosis. J Neurochem 82, 365-374.

Hentati, A., Bejaoui, K., Pericak-Vance, M. A., Hentati, F., Speer, M. C., Hung, W. Y.,Figlewicz, D. A., Haines, J., Rimmler, J., Ben Hamida, C., and et al. (1994). Linkageof recessive familial amyotrophic lateral sclerosis to chromosome 2q33-q35. NatGenet 7, 425-428.

Hentati, A., Ouahchi, K., Pericak-Vance, M. A., Nijhawan, D., Ahmad, A., Yang, Y.,Rimmler, J., Hung, W., Schlotter, B., Ahmed, A., et al. (1998). Linkage of a

Page 97: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

97

commoner form of recessive amyotrophic lateral sclerosis to chromosome 15q15-q22markers. Neurogenetics 2, 55-60.

Hirano, A. (1991). Cytopathology of amyotrophic lateral sclerosis. Adv Neurol 56, 91-101.Hirano, A., Donnenfeld, H., Sasaki, S., and Nakano, I. (1984a). Fine structural observations

of neurofilamentous changes in amyotrophic lateral sclerosis. J Neuropathol ExpNeurol 43, 461-470.

Hirano, A., Nakano, I., Kurland, L. T., Mulder, D. W., Holley, P. W., and Saccomanno, G.(1984b). Fine structural study of neurofibrillary changes in a family with amyotrophiclateral sclerosis. J Neuropathol Exp Neurol 43, 471-480.

Hosler, B. A., Siddique, T., Sapp, P. C., Sailor, W., Huang, M. C., Hossain, A., Daube, J. R.,Nance, M., Fan, C., Kaplan, J., et al. (2000). Linkage of familial amyotrophic lateralsclerosis with frontotemporal dementia to chromosome 9q21-q22. Jama 284, 1664-1669.

Howland, D. S., Liu, J., She, Y., Goad, B., Maragakis, N. J., Kim, B., Erickson, J., Kulik, J.,DeVito, L., Psaltis, G., et al. (2002). Focal loss of the glutamate transporter EAAT2 ina transgenic rat model of SOD1 mutant-mediated amyotrophic lateral sclerosis (ALS).Proc Natl Acad Sci U S A 99, 1604-1609.

Hutton, M., Lendon, C. L., Rizzu, P., Baker, M., Froelich, S., Houlden, H., Pickering-Brown,S., Chakraverty, S., Isaacs, A., Grover, A., et al. (1998). Association of missense and5'-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 393, 702-705.

Inarrea, P., Moini, H., Han, D., Rettori, D., Aguilo, I., Alava, M. A., Iturralde, M., andCadenas, E. (2007). Mitochondrial respiratory chain and thioredoxin reductaseregulate intermembrane Cu,Zn-superoxide dismutase activity: implications formitochondrial energy metabolism and apoptosis. Biochem J 405, 173-179.

Inarrea, P., Moini, H., Rettori, D., Han, D., Martinez, J., Garcia, I., Fernandez-Vizarra, E.,Iturralde, M., and Cadenas, E. (2005). Redox activation of mitochondrialintermembrane space Cu,Zn-superoxide dismutase. Biochem J 387, 203-209.

Jaarsma, D., Haasdijk, E. D., Grashorn, J. A., Hawkins, R., van Duijn, W., Verspaget, H. W.,London, J., and Holstege, J. C. (2000). Human Cu/Zn superoxide dismutase (SOD1)overexpression in mice causes mitochondrial vacuolization, axonal degeneration, andpremature motoneuron death and accelerates motoneuron disease in mice expressing afamilial amyotrophic lateral sclerosis mutant SOD1. Neurobiol Dis 7, 623-643.

Jaarsma, D., Teuling, E., Haasdijk, E. D., De Zeeuw, C. I., and Hoogenraad, C. C. (2008).Neuron-specific expression of mutant superoxide dismutase is sufficient to induceamyotrophic lateral sclerosis in transgenic mice. J Neurosci 28, 2075-2088.

Johnston, J. A., Dalton, M. J., Gurney, M. E., and Kopito, R. R. (2000). Formation of highmolecular weight complexes of mutant Cu, Zn-superoxide dismutase in a mousemodel for familial amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A 97, 12571-12576.

Jonsson, P. A., Ernhill, K., Andersen, P. M., Bergemalm, D., Brannstrom, T., Gredal, O.,Nilsson, P., and Marklund, S. L. (2004). Minute quantities of misfolded mutantsuperoxide dismutase-1 cause amyotrophic lateral sclerosis. Brain 127, 73-88.

Kabashi, E., Agar, J. N., Taylor, D. M., Minotti, S., and Durham, H. D. (2004). Focaldysfunction of the proteasome: a pathogenic factor in a mouse model of amyotrophiclateral sclerosis. J Neurochem 89, 1325-1335.

Kagan, V. E., Tyurin, V. A., Jiang, J., Tyurina, Y. Y., Ritov, V. B., Amoscato, A. A., Osipov,A. N., Belikova, N. A., Kapralov, A. A., Kini, V., et al. (2005). Cytochrome c acts asa cardiolipin oxygenase required for release of proapoptotic factors. Nat Chem Biol 1,223-232.

Kandel, E. R., Schwartz, J. H., and Jessel, T. M. (1991). Diseases of the Motor Unit. InPrinciples of Neural Sciences (Norwalk, Appleton & Lange), pp. 248-250.

Page 98: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

98

Kanekura, K., Hashimoto, Y., Niikura, T., Aiso, S., Matsuoka, M., and Nishimoto, I. (2004).Alsin, the product of ALS2 gene, suppresses SOD1 mutant neurotoxicity throughRhoGEF domain by interacting with SOD1 mutants. J Biol Chem 279, 19247-19256.

Kaspar, B. K., Llado, J., Sherkat, N., Rothstein, J. D., and Gage, F. H. (2003). Retrogradeviral delivery of IGF-1 prolongs survival in a mouse ALS model. Science 301, 839-842.

Kawamata, T., Akiyama, H., Yamada, T., and McGeer, P. L. (1992). Immunologic reactionsin amyotrophic lateral sclerosis brain and spinal cord tissue. Am J Pathol 140, 691-707.

Kiaei, M., Bush, A. I., Morrison, B. M., Morrison, J. H., Cherny, R. A., Volitakis, I., Beal, M.F., and Gordon, J. W. (2004). Genetically decreased spinal cord copper concentrationprolongs life in a transgenic mouse model of amyotrophic lateral sclerosis. J Neurosci24, 7945-7950.

Kiaei, M., Petri, S., Kipiani, K., Gardian, G., Choi, D. K., Chen, J., Calingasan, N. Y.,Schafer, P., Muller, G. W., Stewart, C., et al. (2006). Thalidomide and lenalidomideextend survival in a transgenic mouse model of amyotrophic lateral sclerosis. JNeurosci 26, 2467-2473.

Kim, I., Kim, C. H., Kim, J. H., Lee, J., Choi, J. J., Chen, Z. A., Lee, M. G., Chung, K. C.,Hsu, C. Y., and Ahn, Y. S. (2004). Pyrrolidine dithiocarbamate and zinc inhibitproteasome-dependent proteolysis. Exp Cell Res 298, 229-238.

Kitamei, H., Iwabuchi, K., Namba, K., Yoshida, K., Yanagawa, Y., Kitaichi, N., Kitamura,M., Ohno, S., and Onoe, K. (2006). Amelioration of experimental autoimmuneuveoretinitis (EAU) with an inhibitor of nuclear factor-{kappa}B (NF-{kappa}B),pyrrolidine dithiocarbamate. J Leukoc Biol.

Klivenyi, P., Ferrante, R. J., Matthews, R. T., Bogdanov, M. B., Klein, A. M., Andreassen, O.A., Mueller, G., Wermer, M., Kaddurah-Daouk, R., and Beal, M. F. (1999).Neuroprotective effects of creatine in a transgenic animal model of amyotrophiclateral sclerosis. Nat Med 5, 347-350.

Koh, S. H., Lee, Y. B., Kim, K. S., Kim, H. J., Kim, M., Lee, Y. J., Kim, J., Lee, K. W., andKim, S. H. (2005). Role of GSK-3beta activity in motor neuronal cell death inducedby G93A or A4V mutant hSOD1 gene. Eur J Neurosci 22, 301-309.

Kreutzberg, G. W. (1996). Microglia: a sensor for pathological events in the CNS. TrendsNeurosci 19, 312-318.

Kriz, J., Nguyen, M. D., and Julien, J. P. (2002). Minocycline slows disease progression in amouse model of amyotrophic lateral sclerosis. Neurobiol Dis 10, 268-278.

Kunita, R., Otomo, A., Mizumura, H., Suzuki, K., Showguchi-Miyata, J., Yanagisawa, Y.,Hadano, S., and Ikeda, J. E. (2004). Homo-oligomerization of ALS2 through itsunique carboxyl-terminal regions is essential for the ALS2-associated Rab5 guaninenucleotide exchange activity and its regulatory function on endosome trafficking. JBiol Chem 279, 38626-38635.

Kunst, C. B., Messer, L., Gordon, J., Haines, J., and Patterson, D. (2000). Genetic mapping ofa mouse modifier gene that can prevent ALS onset. Genomics 70, 181-189.

Kurland, L. T., and Mulder, D. W. (1955). Epidemiologic investigations of amyotrophiclateral sclerosis. 2. Familial aggregations indicative of dominant inheritance. I.Neurology 5, 182-196.

La Rosa, G., Cardali, S., Genovese, T., Conti, A., Di Paola, R., La Torre, D., Cacciola, F., andCuzzocrea, S. (2004). Inhibition of the nuclear factor-kappaB activation withpyrrolidine dithiocarbamate attenuating inflammation and oxidative stress afterexperimental spinal cord trauma in rats. J Neurosurg Spine 1, 311-321.

Lacomblez, L., Bensimon, G., Leigh, P. N., Guillet, P., and Meininger, V. (1996). Dose-ranging study of riluzole in amyotrophic lateral sclerosis. Amyotrophic LateralSclerosis/Riluzole Study Group II. Lancet 347, 1425-1431.

Page 99: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

99

Lai, E. C., Felice, K. J., Festoff, B. W., Gawel, M. J., Gelinas, D. F., Kratz, R., Murphy, M.F., Natter, H. M., Norris, F. H., and Rudnicki, S. A. (1997). Effect of recombinanthuman insulin-like growth factor-I on progression of ALS. A placebo-controlledstudy. The North America ALS/IGF-I Study Group. Neurology 49, 1621-1630.

Lambrechts, D., Storkebaum, E., Morimoto, M., Del-Favero, J., Desmet, F., Marklund, S. L.,Wyns, S., Thijs, V., Andersson, J., van Marion, I., et al. (2003). VEGF is a modifier ofamyotrophic lateral sclerosis in mice and humans and protects motoneurons againstischemic death. Nat Genet 34, 383-394.

Lawrence, A., Jones, C. M., Wardman, P., and Burkitt, M. J. (2003). Evidence for the role ofa peroxidase compound I-type intermediate in the oxidation of glutathione, NADH,ascorbate, and dichlorofluorescin by cytochrome c/H2O2. Implications for oxidativestress during apoptosis. J Biol Chem 278, 29410-29419.

Lee, M. K., Marszalek, J. R., and Cleveland, D. W. (1994). A mutant neurofilament subunitcauses massive, selective motor neuron death: implications for the pathogenesis ofhuman motor neuron disease. Neuron 13, 975-988.

Lei, X. G., Zhu, J. H., McClung, J. P., Aregullin, M., and Roneker, C. A. (2006). Micedeficient in Cu,Zn-superoxide dismutase are resistant to acetaminophen toxicity.Biochem J 399, 455-461.

Li, M., Ona, V. O., Guegan, C., Chen, M., Jackson-Lewis, V., Andrews, L. J., Olszewski, A.J., Stieg, P. E., Lee, J. P., Przedborski, S., and Friedlander, R. M. (2000). Functionalrole of caspase-1 and caspase-3 in an ALS transgenic mouse model. Science 288, 335-339.

Lino, M. M., Schneider, C., and Caroni, P. (2002). Accumulation of SOD1 mutants inpostnatal motoneurons does not cause motoneuron pathology or motoneuron disease. JNeurosci 22, 4825-4832.

Liochev, S. I., and Fridovich, I. (2000). Copper- and zinc-containing superoxide dismutasecan act as a superoxide reductase and a superoxide oxidase. J Biol Chem 275, 38482-38485.

Liu, J., Lillo, C., Jonsson, P. A., Vande Velde, C., Ward, C. M., Miller, T. M., Subramaniam,J. R., Rothstein, J. D., Marklund, S., Andersen, P. M., et al. (2004). Toxicity offamilial ALS-linked SOD1 mutants from selective recruitment to spinal mitochondria.Neuron 43, 5-17.

Liu, J., Shinobu, L. A., Ward, C. M., Young, D., and Cleveland, D. W. (2005). Elevation ofthe Hsp70 chaperone does not effect toxicity in mouse models of familial amyotrophiclateral sclerosis. J Neurochem 93, 875-882.

Liu, S. F., Ye, X., and Malik, A. B. (1999). Inhibition of NF-kappaB activation by pyrrolidinedithiocarbamate prevents In vivo expression of proinflammatory genes. Circulation100, 1330-1337.

Lobsiger, C. S., Garcia, M. L., Ward, C. M., and Cleveland, D. W. (2005). Altered axonalarchitecture by removal of the heavily phosphorylated neurofilament tail domainsstrongly slows superoxide dismutase 1 mutant-mediated ALS. Proc Natl Acad Sci U SA 102, 10351-10356.

Malm, T. M., Iivonen, H., Goldsteins, G., Keksa-Goldsteine, V., Ahtoniemi, T., Kanninen,K., Salminen, A., Auriola, S., Van Groen, T., Tanila, H., and Koistinaho, J. (2007).Pyrrolidine dithiocarbamate activates Akt and improves spatial learning in APP/PS1mice without affecting beta-amyloid burden. J Neurosci 27, 3712-3721.

Martin, H., Eckerskorn, C., Gartner, F., Rassow, J., Lottspeich, F., and Pfanner, N. (1998).The yeast mitochondrial intermembrane space: purification and analysis of twodistinct fractions. Anal Biochem 265, 123-128.

Martin, N., Jaubert, J., Gounon, P., Salido, E., Haase, G., Szatanik, M., and Guenet, J. L.(2002). A missense mutation in Tbce causes progressive motor neuronopathy in mice.Nat Genet 32, 443-447.

Page 100: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

100

Martinou, J. C., Martinou, I., and Kato, A. C. (1992). Cholinergic differentiation factor(CDF/LIF) promotes survival of isolated rat embryonic motoneurons in vitro. Neuron8, 737-744.

Matsui, N., Kasajima, K., Hada, M., Nagata, T., Senga, N., Yasui, Y., Fukuishi, N., andAkagi, M. (2005). Inhibiton of NF-kappaB activation during ischemia reduces hepaticischemia/reperfusion injury in rats. J Toxicol Sci 30, 103-110.

Mattiazzi, M., D'Aurelio, M., Gajewski, C. D., Martushova, K., Kiaei, M., Beal, M. F., andManfredi, G. (2002). Mutated human SOD1 causes dysfunction of oxidativephosphorylation in mitochondria of transgenic mice. J Biol Chem 277, 29626-29633.

Mattson, M. P., and Camandola, S. (2001). NF-kappaB in neuronal plasticity andneurodegenerative disorders. J Clin Invest 107, 247-254.

McGeer, E. G., and McGeer, P. L. (1999). Brain inflammation in Alzheimer disease and thetherapeutic implications. Curr Pharm Des 5, 821-836.

McGeer, E. G., and McGeer, P. L. (2005). Pharmacologic approaches to the treatment ofamyotrophic lateral sclerosis. BioDrugs 19, 31-37.

McGeer, P. L., McGeer, E. G., Kawamata, T., Yamada, T., and Akiyama, H. (1991).Reactions of the immune system in chronic degenerative neurological diseases. Can JNeurol Sci 18, 376-379.

Messer, A., and Flaherty, L. (1986). Autosomal dominance in a late-onset motor neurondisease in the mouse. J Neurogenet 3, 345-355.

Messer, A., Plummer, J., Wong, V., and Lavail, M. M. (1993). Retinal degeneration in motorneuron degeneration (mnd) mutant mice. Exp Eye Res 57, 637-641.

Messer, A., Strominger, N. L., and Mazurkiewicz, J. E. (1987). Histopathology of the late-onset motor neuron degeneration (Mnd) mutant in the mouse. J Neurogenet 4, 201-213.

Messina, S., Bitto, A., Aguennouz, M., Minutoli, L., Monici, M. C., Altavilla, D., Squadrito,F., and Vita, G. (2006). Nuclear factor kappa-B blockade reduces skeletal muscledegeneration and enhances muscle function in Mdx mice. Exp Neurol 198, 234-241.

Migheli, A., Piva, R., Atzori, C., Troost, D., and Schiffer, D. (1997). c-Jun, JNK/SAPKkinases and transcription factor NF-kappa B are selectively activated in astrocytes, butnot motor neurons, in amyotrophic lateral sclerosis. J Neuropathol Exp Neurol 56,1314-1322.

Miller, T. M., Kaspar, B. K., Kops, G. J., Yamanaka, K., Christian, L. J., Gage, F. H., andCleveland, D. W. (2005). Virus-delivered small RNA silencing sustains strength inamyotrophic lateral sclerosis. Ann Neurol 57, 773-776.

Mitsumoto, H., and Bradley, W. G. (1982). Murine motor neuron disease (the wobblermouse): degeneration and regeneration of the lower motor neuron. Brain 105 (Pt 4),811-834.

Moreira, M. C., Klur, S., Watanabe, M., Nemeth, A. H., Le Ber, I., Moniz, J. C., Tranchant,C., Aubourg, P., Tazir, M., Schols, L., et al. (2004). Senataxin, the ortholog of a yeastRNA helicase, is mutant in ataxia-ocular apraxia 2. Nat Genet 36, 225-227.

Nagai, M., Aoki, M., Miyoshi, I., Kato, M., Pasinelli, P., Kasai, N., Brown, R. H., Jr., andItoyama, Y. (2001). Rats expressing human cytosolic copper-zinc superoxidedismutase transgenes with amyotrophic lateral sclerosis: associated mutations developmotor neuron disease. J Neurosci 21, 9246-9254.

Nagai, M., Re, D. B., Nagata, T., Chalazonitis, A., Jessell, T. M., Wichterle, H., andPrzedborski, S. (2007). Astrocytes expressing ALS-linked mutated SOD1 releasefactors selectively toxic to motor neurons. Nat Neurosci 10, 615-622.

Nagano, I., Ilieva, H., Shiote, M., Murakami, T., Yokoyama, M., Shoji, M., and Abe, K.(2005a). Therapeutic benefit of intrathecal injection of insulin-like growth factor-1 ina mouse model of Amyotrophic Lateral Sclerosis. J Neurol Sci 235, 61-68.

Page 101: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

101

Nagano, I., Shiote, M., Murakami, T., Kamada, H., Hamakawa, Y., Matsubara, E.,Yokoyama, M., Moritaz, K., Shoji, M., and Abe, K. (2005b). Beneficial effects ofintrathecal IGF-1 administration in patients with amyotrophic lateral sclerosis. NeurolRes 27, 768-772.

Nagano, I., Wong, P. C., and Rothstein, J. D. (1996). Nitration of glutamate transporters intransgenic mice with a familial amyotrophic lateral sclerosis-linked SOD1 mutation.Ann Neurol 40, 542.

Nguyen, M. D., Julien, J. P., and Rivest, S. (2001). Induction of proinflammatory moleculesin mice with amyotrophic lateral sclerosis: no requirement for proapoptoticinterleukin-1beta in neurodegeneration. Ann Neurol 50, 630-639.

Nishimura, A. L., Mitne-Neto, M., Silva, H. C., Richieri-Costa, A., Middleton, S., Cascio, D.,Kok, F., Oliveira, J. R., Gillingwater, T., Webb, J., et al. (2004). A mutation in thevesicle-trafficking protein VAPB causes late-onset spinal muscular atrophy andamyotrophic lateral sclerosis. Am J Hum Genet 75, 822-831.

Nurmi, A., Goldsteins, G., Narvainen, J., Pihlaja, R., Ahtoniemi, T., Grohn, O., andKoistinaho, J. (2006). Antioxidant pyrrolidine dithiocarbamate activates Akt-GSKsignaling and is neuroprotective in neonatal hypoxia-ischemia. Free Radic Biol Med40, 1776-1784.

Nurmi, A., Lindsberg, P. J., Koistinaho, M., Zhang, W., Juettler, E., Karjalainen-Lindsberg,M. L., Weih, F., Frank, N., Schwaninger, M., and Koistinaho, J. (2004a). Nuclearfactor-kappaB contributes to infarction after permanent focal ischemia. Stroke 35,987-991.

Nurmi, A., Vartiainen, N., Pihlaja, R., Goldsteins, G., Yrjanheikki, J., and Koistinaho, J.(2004b). Pyrrolidine dithiocarbamate inhibits translocation of nuclear factor kappa-Bin neurons and protects against brain ischaemia with a wide therapeutic time window.J Neurochem 91, 755-765.

Oberley, L. W., and Spitz, D. R. (1984). Assay of superoxide dismutase activity in tumortissue. Methods Enzymol 105, 457-464.

Oeda, T., Shimohama, S., Kitagawa, N., Kohno, R., Imura, T., Shibasaki, H., and Ishii, N.(2001). Oxidative stress causes abnormal accumulation of familial amyotrophic lateralsclerosis-related mutant SOD1 in transgenic Caenorhabditis elegans. Hum Mol Genet10, 2013-2023.

Okado-Matsumoto, A., and Fridovich, I. (2001). Subcellular distribution of superoxidedismutases (SOD) in rat liver: Cu,Zn-SOD in mitochondria. J Biol Chem 276, 38388-38393.

Oosthuyse, B., Moons, L., Storkebaum, E., Beck, H., Nuyens, D., Brusselmans, K., VanDorpe, J., Hellings, P., Gorselink, M., Heymans, S., et al. (2001). Deletion of thehypoxia-response element in the vascular endothelial growth factor promoter causesmotor neuron degeneration. Nat Genet 28, 131-138.

Orrell, R., de Belleroche, J., Marklund, S., Bowe, F., and Hallewell, R. (1995). A novel SODmutant and ALS. Nature 374, 504-505.

Otomo, A., Hadano, S., Okada, T., Mizumura, H., Kunita, R., Nishijima, H., Showguchi-Miyata, J., Yanagisawa, Y., Kohiki, E., Suga, E., et al. (2003). ALS2, a novel guaninenucleotide exchange factor for the small GTPase Rab5, is implicated in endosomaldynamics. Hum Mol Genet 12, 1671-1687.

Parkes, T. L., Elia, A. J., Dickinson, D., Hilliker, A. J., Phillips, J. P., and Boulianne, G. L.(1998). Extension of Drosophila lifespan by overexpression of human SOD1 inmotorneurons. Nat Genet 19, 171-174.

Parodi, F. E., Mao, D., Ennis, T. L., Bartoli, M. A., and Thompson, R. W. (2005).Suppression of experimental abdominal aortic aneurysms in mice by treatment withpyrrolidine dithiocarbamate, an antioxidant inhibitor of nuclear factor-kappaB. J VascSurg 41, 479-489.

Page 102: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

102

Parton, M. J., Broom, W., Andersen, P. M., Al-Chalabi, A., Nigel Leigh, P., Powell, J. F., andShaw, C. E. (2002). D90A-SOD1 mediated amyotrophic lateral sclerosis: a singlefounder for all cases with evidence for a Cis-acting disease modifier in the recessivehaplotype. Hum Mutat 20, 473.

Pasinelli, P., Belford, M. E., Lennon, N., Bacskai, B. J., Hyman, B. T., Trotti, D., and Brown,R. H., Jr. (2004). Amyotrophic lateral sclerosis-associated SOD1 mutant proteins bindand aggregate with Bcl-2 in spinal cord mitochondria. Neuron 43, 19-30.

Pasinelli, P., Houseweart, M. K., Brown, R. H., Jr., and Cleveland, D. W. (2000). Caspase-1and -3 are sequentially activated in motor neuron death in Cu,Zn superoxidedismutase-mediated familial amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A97, 13901-13906.

Patel, Y. J., Payne Smith, M. D., de Belleroche, J., and Latchman, D. S. (2005). Hsp27 andHsp70 administered in combination have a potent protective effect against FALS-associated SOD1-mutant-induced cell death in mammalian neuronal cells. Brain ResMol Brain Res 134, 256-274.

Pereverzev, M. O., Vygodina, T. V., Konstantinov, A. A., and Skulachev, V. P. (2003).Cytochrome c, an ideal antioxidant. Biochem Soc Trans 31, 1312-1315.

Pramatarova, A., Laganiere, J., Roussel, J., Brisebois, K., and Rouleau, G. A. (2001). Neuron-specific expression of mutant superoxide dismutase 1 in transgenic mice does not leadto motor impairment. J Neurosci 21, 3369-3374.

Pringle, C. E., Hudson, A. J., Munoz, D. G., Kiernan, J. A., Brown, W. F., and Ebers, G. C.(1992). Primary lateral sclerosis. Clinical features, neuropathology and diagnosticcriteria. Brain 115 ( Pt 2), 495-520.

Puls, I., Jonnakuty, C., LaMonte, B. H., Holzbaur, E. L., Tokito, M., Mann, E., Floeter, M. K.,Bidus, K., Drayna, D., Oh, S. J., et al. (2003). Mutant dynactin in motor neurondisease. Nat Genet 33, 455-456.

Pun, S., Santos, A. F., Saxena, S., Xu, L., and Caroni, P. (2006). Selective vulnerability andpruning of phasic motoneuron axons in motoneuron disease alleviated by CNTF. NatNeurosci 9, 408-419.

Puttaparthi, K., and Elliott, J. L. (2005). Non-neuronal induction of immunoproteasomesubunits in an ALS model: possible mediation by cytokines. Exp Neurol 196, 441-451.

Rademakers, R., Cruts, M., and van Broeckhoven, C. (2004). The role of tau (MAPT) infrontotemporal dementia and related tauopathies. Hum Mutat 24, 277-295.

Ralph, G. S., Radcliffe, P. A., Day, D. M., Carthy, J. M., Leroux, M. A., Lee, D. C., Wong, L.F., Bilsland, L. G., Greensmith, L., Kingsman, S. M., et al. (2005). Silencing mutantSOD1 using RNAi protects against neurodegeneration and extends survival in an ALSmodel. Nat Med 11, 429-433.

Ranta, S., Zhang, Y., Ross, B., Lonka, L., Takkunen, E., Messer, A., Sharp, J., Wheeler, R.,Kusumi, K., Mole, S., et al. (1999). The neuronal ceroid lipofuscinoses in humanEPMR and mnd mutant mice are associated with mutations in CLN8. Nat Genet 23,233-236.

Ratovitski, T., Corson, L. B., Strain, J., Wong, P., Cleveland, D. W., Culotta, V. C., andBorchelt, D. R. (1999). Variation in the biochemical/biophysical properties of mutantsuperoxide dismutase 1 enzymes and the rate of disease progression in familialamyotrophic lateral sclerosis kindreds. Hum Mol Genet 8, 1451-1460.

Reaume, A. G., Elliott, J. L., Hoffman, E. K., Kowall, N. W., Ferrante, R. J., Siwek, D. F.,Wilcox, H. M., Flood, D. G., Beal, M. F., Brown, R. H., Jr., et al. (1996). Motorneurons in Cu/Zn superoxide dismutase-deficient mice develop normally but exhibitenhanced cell death after axonal injury. Nat Genet 13, 43-47.

Reisinger, E. C., Kern, P., Ernst, M., Bock, P., Flad, H. D., and Dietrich, M. (1990). Inhibitionof HIV progression by dithiocarb. German DTC Study Group. Lancet 335, 679-682.

Page 103: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

103

Rigobello, M. P., Donella-Deana, A., Cesaro, L., and Bindoli, A. (2001). Distribution ofprotein disulphide isomerase in rat liver mitochondria. Biochem J 356, 567-570.

Ripps, M. E., Huntley, G. W., Hof, P. R., Morrison, J. H., and Gordon, J. W. (1995).Transgenic mice expressing an altered murine superoxide dismutase gene provide ananimal model of amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A 92, 689-693.

Rocha, J. A., Reis, C., Simoes, F., Fonseca, J., and Mendes Ribeiro, J. (2005). Diagnosticinvestigation and multidisciplinary management in motor neuron disease. J Neurol252, 1435-1447.

Rosen, D. R., Siddique, T., Patterson, D., Figlewicz, D. A., Sapp, P., Hentati, A., Donaldson,D., Goto, J., O'Regan, J. P., Deng, H. X., and et al. (1993). Mutations in Cu/Znsuperoxide dismutase gene are associated with familial amyotrophic lateral sclerosis.Nature 362, 59-62.

Rothstein, J. D., Dykes-Hoberg, M., Pardo, C. A., Bristol, L. A., Jin, L., Kuncl, R. W., Kanai,Y., Hediger, M. A., Wang, Y., Schielke, J. P., and Welty, D. F. (1996). Knockout ofglutamate transporters reveals a major role for astroglial transport in excitotoxicity andclearance of glutamate. Neuron 16, 675-686.

Rothstein, J. D., Jin, L., Dykes-Hoberg, M., and Kuncl, R. W. (1993). Chronic inhibition ofglutamate uptake produces a model of slow neurotoxicity. Proc Natl Acad Sci U S A90, 6591-6595.

Rothstein, J. D., Kuncl, R., Chaudhry, V., Clawson, L., Cornblath, D. R., Coyle, J. T., andDrachman, D. B. (1991). Excitatory amino acids in amyotrophic lateral sclerosis: anupdate. Ann Neurol 30, 224-225.

Rothstein, J. D., and Kuncl, R. W. (1995). Neuroprotective strategies in a model of chronicglutamate-mediated motor neuron toxicity. J Neurochem 65, 643-651.

Rothstein, J. D., Patel, S., Regan, M. R., Haenggeli, C., Huang, Y. H., Bergles, D. E., Jin, L.,Dykes Hoberg, M., Vidensky, S., Chung, D. S., et al. (2005). Beta-lactam antibioticsoffer neuroprotection by increasing glutamate transporter expression. Nature 433, 73-77.

Rothstein, J. D., Tsai, G., Kuncl, R. W., Clawson, L., Cornblath, D. R., Drachman, D. B.,Pestronk, A., Stauch, B. L., and Coyle, J. T. (1990). Abnormal excitatory amino acidmetabolism in amyotrophic lateral sclerosis. Ann Neurol 28, 18-25.

Rothstein, J. D., Van 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, 73-84.

Sapp, P. C., Hosler, B. A., McKenna-Yasek, D., Chin, W., Gann, A., Genise, H., Gorenstein,J., Huang, M., Sailer, W., Scheffler, M., et al. (2003). Identification of two novel locifor dominantly inherited familial amyotrophic lateral sclerosis. Am J Hum Genet 73,397-403.

Schaefer, A. M., Sanes, J. R., and Lichtman, J. W. (2005). A compensatory subpopulation ofmotor neurons in a mouse model of amyotrophic lateral sclerosis. J Comp Neurol 490,209-219.

Schmalbruch, H., Jensen, H. J., Bjaerg, M., Kamieniecka, Z., and Kurland, L. (1991). A newmouse mutant with progressive motor neuronopathy. J Neuropathol Exp Neurol 50,192-204.

Schmitt-John, T., Drepper, C., Mussmann, A., Hahn, P., Kuhlmann, M., Thiel, C., Hafner, M.,Lengeling, A., Heimann, P., Jones, J. M., et al. (2005). Mutation of Vps54 causesmotor neuron disease and defective spermiogenesis in the wobbler mouse. Nat Genet37, 1213-1215.

Schreck, R., Meier, B., Mannel, D. N., Droge, W., and Baeuerle, P. A. (1992).Dithiocarbamates as potent inhibitors of nuclear factor kappa B activation in intactcells. J Exp Med 175, 1181-1194.

Page 104: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

104

Scott, S., Kranz, J. E., Cole, J., Lincecum, J. M., Thompson, K., Kelly, N., Bostrom, A.,Theodoss, J., Al-Nakhala, B. M., Vieira, F. G., et al. (2008). Design, power, andinterpretation of studies in the standard murine model of ALS. Amyotroph LateralScler 9, 4-15.

Shaw, P. J., Forrest, V., Ince, P. G., Richardson, J. P., and Wastell, H. J. (1995a). CSF andplasma amino acid levels in motor neuron disease: elevation of CSF glutamate in asubset of patients. Neurodegeneration 4, 209-216.

Shaw, P. J., Ince, P. G., Falkous, G., and Mantle, D. (1995b). Oxidative damage to protein insporadic motor neuron disease spinal cord. Ann Neurol 38, 691-695.

Shefner, J. M., Cudkowicz, M. E., Schoenfeld, D., Conrad, T., Taft, J., Chilton, M., Urbinelli,L., Qureshi, M., Zhang, H., Pestronk, A., et al. (2004). A clinical trial of creatine inALS. Neurology 63, 1656-1661.

Simpson, C. L., and Al-Chalabi, A. (2006). Amyotrophic lateral sclerosis as a complexgenetic disease. Biochim Biophys Acta 1762, 973-985.

Singh, R. J., Karoui, H., Gunther, M. R., Beckman, J. S., Mason, R. P., and Kalyanaraman, B.(1998). Reexamination of the mechanism of hydroxyl radical adducts formed from thereaction between familial amyotrophic lateral sclerosis-associated Cu,Zn superoxidedismutase mutants and H2O2. Proc Natl Acad Sci U S A 95, 6675-6680.

Sjalander, A., Beckman, G., Deng, H. X., Iqbal, Z., Tainer, J. A., and Siddique, T. (1995).The D90A mutation results in a polymorphism of Cu,Zn superoxide dismutase that isprevalent in northern Sweden and Finland. Hum Mol Genet 4, 1105-1108.

Smith, R. A., Miller, T. M., Yamanaka, K., Monia, B. P., Condon, T. P., Hung, G., Lobsiger,C. S., Ward, C. M., McAlonis-Downes, M., Wei, H., et al. (2006). Antisenseoligonucleotide therapy for neurodegenerative disease. J Clin Invest 116, 2290-2296.

Solomon, J. N., Lewis, C. A., Ajami, B., Corbel, S. Y., Rossi, F. M., and Krieger, C. (2006).Origin and distribution of bone marrow-derived cells in the central nervous system ina mouse model of amyotrophic lateral sclerosis. Glia 53, 744-753.

Spreux-Varoquaux, O., Bensimon, G., Lacomblez, L., Salachas, F., Pradat, P. F., Le Forestier,N., Marouan, A., Dib, M., and Meininger, V. (2002). Glutamate levels incerebrospinal fluid in amyotrophic lateral sclerosis: a reappraisal using a new HPLCmethod with coulometric detection in a large cohort of patients. J Neurol Sci 193, 73-78.

Squire, L. R., Bloom, F. E., McConnell, S. K., Roberts, J. L., Spitzer, N. C., and Zigmond, M.J. (2003a). Fundamentals of Motor Systems. In Fundamental Neuroscience (SanDiego, Academic Press), pp. 755-765.

Squire, L. R., Bloom, F. E., McConnell, S. K., Roberts, J. L., Spitzer, N. C., and Zigmond, M.J. (2003b). The Spinal Cord, Muscle and Locomotion. In Fundamental Neuroscience(San Diego, Academic Press), pp. 767-789.

Sreedharan, J., Blair, I. P., Tripathi, V. B., Hu, X., Vance, C., Rogelj, B., Ackerley, S.,Durnall, J. C., Williams, K. L., Buratti, E., et al. (2008). TDP-43 Mutations inFamilial and Sporadic Amyotrophic Lateral Sclerosis. Science.

Stohwasser, R., Giesebrecht, J., Kraft, R., Muller, E. C., Hausler, K. G., Kettenmann, H.,Hanisch, U. K., and Kloetzel, P. M. (2000). Biochemical analysis of proteasomes frommouse microglia: induction of immunoproteasomes by interferon-gamma andlipopolysaccharide. Glia 29, 355-365.

Storkebaum, E., Lambrechts, D., Dewerchin, M., Moreno-Murciano, M. P., Appelmans, S.,Oh, H., Van Damme, P., Rutten, B., Man, W. Y., De Mol, M., et al. (2005). Treatmentof motoneuron degeneration by intracerebroventricular delivery of VEGF in a ratmodel of ALS. Nat Neurosci 8, 85-92.

Sturtz, L. A., Diekert, K., Jensen, L. T., Lill, R., and Culotta, V. C. (2001). A fraction of yeastCu,Zn-superoxide dismutase and its metallochaperone, CCS, localize to the

Page 105: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

105

intermembrane space of mitochondria. A physiological role for SOD1 in guardingagainst mitochondrial oxidative damage. J Biol Chem 276, 38084-38089.

Takeuchi, H., Kobayashi, Y., Yoshihara, T., Niwa, J., Doyu, M., Ohtsuka, K., and Sobue, G.(2002). Hsp70 and Hsp40 improve neurite outgrowth and suppress intracytoplasmicaggregate formation in cultured neuronal cells expressing mutant SOD1. Brain Res949, 11-22.

Tanaka, K., Watase, K., Manabe, T., Yamada, K., Watanabe, M., Takahashi, K., Iwama, H.,Nishikawa, T., Ichihara, N., Kikuchi, T., et al. (1997). Epilepsy and exacerbation ofbrain injury in mice lacking the glutamate transporter GLT-1. Science 276, 1699-1702.

Terry, P. D., Kamel, F., Umbach, D. M., Lehman, T. A., Hu, H., Sandler, D. P., and Taylor, J.A. (2004). VEGF promoter haplotype and amyotrophic lateral sclerosis (ALS). JNeurogenet 18, 429-434.

Troost, D., Claessen, N., van den Oord, J. J., Swaab, D. F., and de Jong, J. M. (1993).Neuronophagia in the motor cortex in amyotrophic lateral sclerosis. Neuropathol ApplNeurobiol 19, 390-397.

Tummala, H., Jung, C., Tiwari, A., Higgins, C. M., Hayward, L. J., and Xu, Z. (2005).Inhibition of chaperone activity is a shared property of several Cu,Zn-superoxidedismutase mutants that cause amyotrophic lateral sclerosis. J Biol Chem 280, 17725-17731.

Turano, C., Coppari, S., Altieri, F., and Ferraro, A. (2002). Proteins of the PDI family:unpredicted non-ER locations and functions. J Cell Physiol 193, 154-163.

Turner, M. R., Cagnin, A., Turkheimer, F. E., Miller, C. C., Shaw, C. E., Brooks, D. J., Leigh,P. N., and Banati, R. B. (2004). Evidence of widespread cerebral microglial activationin amyotrophic lateral sclerosis: an [11C](R)-PK11195 positron emission tomographystudy. Neurobiol Dis 15, 601-609.

Urushitani, M., Ezzi, S. A., and Julien, J. P. (2007). Therapeutic effects of immunization withmutant superoxide dismutase in mice models of amyotrophic lateral sclerosis. ProcNatl Acad Sci U S A 104, 2495-2500.

Urushitani, M., Sik, A., Sakurai, T., Nukina, N., Takahashi, R., and Julien, J. P. (2006).Chromogranin-mediated secretion of mutant superoxide dismutase proteins linked toamyotrophic lateral sclerosis. Nat Neurosci 9, 108-118.

Valentine, J. S., Doucette, P. A., and Zittin Potter, S. (2005). Copper-zinc superoxidedismutase and amyotrophic lateral sclerosis. Annu Rev Biochem 74, 563-593.

Van Damme, P., Van Den Bosch, L., Van Houtte, E., Callewaert, G., and Robberecht, W.(2002). GluR2-dependent properties of AMPA receptors determine the selectivevulnerability of motor neurons to excitotoxicity. J Neurophysiol 88, 1279-1287.

Van Den Bosch, L., Tilkin, P., Lemmens, G., and Robberecht, W. (2002). Minocycline delaysdisease onset and mortality in a transgenic model of ALS. Neuroreport 13, 1067-1070.

Vijayvergiya, C., Beal, M. F., Buck, J., and Manfredi, G. (2005). Mutant superoxidedismutase 1 forms aggregates in the brain mitochondrial matrix of amyotrophic lateralsclerosis mice. J Neurosci 25, 2463-2470.

Vleminckx, V., Van Damme, P., Goffin, K., Delye, H., Van Den Bosch, L., and Robberecht,W. (2002). Upregulation of HSP27 in a transgenic model of ALS. J Neuropathol ExpNeurol 61, 968-974.

Wang, J., Slunt, H., Gonzales, V., Fromholt, D., Coonfield, M., Copeland, N. G., Jenkins, N.A., and Borchelt, D. R. (2003). Copper-binding-site-null SOD1 causes ALS intransgenic mice: aggregates of non-native SOD1 delineate a common feature. HumMol Genet 12, 2753-2764.

Wang, J., Xu, G., Gonzales, V., Coonfield, M., Fromholt, D., Copeland, N. G., Jenkins, N. A.,and Borchelt, D. R. (2002). Fibrillar inclusions and motor neuron degeneration in

Page 106: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

106

transgenic mice expressing superoxide dismutase 1 with a disrupted copper-bindingsite. Neurobiol Dis 10, 128-138.

Wang, J., Xu, G., Slunt, H. H., Gonzales, V., Coonfield, M., Fromholt, D., Copeland, N. G.,Jenkins, N. A., and Borchelt, D. R. (2005). Coincident thresholds of mutant protein forparalytic disease and protein aggregation caused by restrictively expressed superoxidedismutase cDNA. Neurobiol Dis 20, 943-952.

Watanabe, M., Dykes-Hoberg, M., Culotta, V. C., Price, D. L., Wong, P. C., and Rothstein, J.D. (2001). Histological evidence of protein aggregation in mutant SOD1 transgenicmice and in amyotrophic lateral sclerosis neural tissues. Neurobiol Dis 8, 933-941.

Wiedau-Pazos, M., Goto, J. J., Rabizadeh, S., Gralla, E. B., Roe, J. A., Lee, M. K., Valentine,J. S., and Bredesen, D. E. (1996). Altered reactivity of superoxide dismutase infamilial amyotrophic lateral sclerosis. Science 271, 515-518.

Williamson, T. L., Bruijn, L. I., Zhu, Q., Anderson, K. L., Anderson, S. D., Julien, J. P., andCleveland, D. W. (1998). Absence of neurofilaments reduces the selectivevulnerability of motor neurons and slows disease caused by a familial amyotrophiclateral sclerosis-linked superoxide dismutase 1 mutant. Proc Natl Acad Sci U S A 95,9631-9636.

Williamson, T. L., and Cleveland, D. W. (1999). Slowing of axonal transport is a very earlyevent in the toxicity of ALS-linked SOD1 mutants to motor neurons. Nat Neurosci 2,50-56.

Williamson, T. L., Corson, L. B., Huang, L., Burlingame, A., Liu, J., Bruijn, L. I., andCleveland, D. W. (2000). Toxicity of ALS-linked SOD1 mutants. Science 288, 399.

Witan, H., Kern, A., Koziollek-Drechsler, I., Wade, R., Behl, C., and Clement, A. M. (2008).Heterodimer formation of wild-type and amyotrophic lateral sclerosis-causing mutantCu/Zn-Superoxide dismutase induces toxicity independent of protein aggregation.Hum Mol Genet.

Wolvetang, E. J., Johnson, K. L., Krauer, K., Ralph, S. J., and Linnane, A. W. (1994).Mitochondrial respiratory chain inhibitors induce apoptosis. FEBS Lett 339, 40-44.

Wong, P. C., Pardo, C. A., Borchelt, D. R., Lee, M. K., Copeland, N. G., Jenkins, N. A.,Sisodia, S. S., Cleveland, D. W., and Price, D. L. (1995). An adverse property of afamilial ALS-linked SOD1 mutation causes motor neuron disease characterized byvacuolar degeneration of mitochondria. Neuron 14, 1105-1116.

Wu, A. S., Kiaei, M., Aguirre, N., Crow, J. P., Calingasan, N. Y., Browne, S. E., and Beal, M.F. (2003). Iron porphyrin treatment extends survival in a transgenic animal model ofamyotrophic lateral sclerosis. J Neurochem 85, 142-150.

Ying, W., Anderson, C. M., Chen, Y., Stein, B. A., Fahlman, C. S., Copin, J. C., Chan, P. H.,and Swanson, R. A. (2000). Differing effects of copper,zinc superoxide dismutaseoverexpression on neurotoxicity elicited by nitric oxide, reactive oxygen species, andexcitotoxins. J Cereb Blood Flow Metab 20, 359-368.

Yoo, H. Y., Kim, S. S., and Rho, H. M. (1999). Overexpression and simple purification ofhuman superoxide dismutase (SOD1) in yeast and its resistance to oxidative stress. JBiotechnol 68, 29-35.

Yoshida, S., Mulder, D. W., Kurland, L. T., Chu, C. P., and Okazaki, H. (1986). Follow-upstudy on amyotrophic lateral sclerosis in Rochester, Minn., 1925 through 1984.Neuroepidemiology 5, 61-70.

Yrjänheikki, J., Tikka, T., Keinänen, R., Goldsteins, G., Chan, P. H., and Koistinaho, J.(1999). A tetracycline derivative, minocycline, reduces inflammation and protectsagainst focal cerebral ischemia with a wide therapeutic window. Proc Natl Acad Sci US A 96, 13496-13500.

Zelko, I. N., Mariani, T. J., and Folz, R. J. (2002). Superoxide dismutase multigene family: acomparison of the CuZn-SOD (SOD1), Mn-SOD (SOD2), and EC-SOD (SOD3) genestructures, evolution, and expression. Free Radic Biol Med 33, 337-349.

Page 107: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

107

Zhu, S., Stavrovskaya, I. G., Drozda, M., Kim, B. Y., Ona, V., Li, M., Sarang, S., Liu, A. S.,Hartley, D. M., Wu du, C., et al. (2002). Minocycline inhibits cytochrome c releaseand delays progression of amyotrophic lateral sclerosis in mice. Nature 417, 74-78.

Page 108: TONI AHTONIEMI Mutant Cu,Zn Superoxide Dismutase in ... · the selective destruction of upper and lower motor neurons in the spinal cord, ... Professor Dan Lindholm, ... 2.1.1.2 Motor

Kuopio University Publications G. - A.I.Virtanen Institute G 43. Nairismägi, Jaak. Magnetic resonance imaging study of induced epileptogenesis in animal models of epilepsy. 2006. 77 p. Acad. Diss. G 44. Niiranen, Kirsi. Consequences of spermine synthase or spermidine/spermine N1-acetyltransferase deficiency in polyamine metabolism - Studies with gene-disrupted embryonic stem cells and mice. 2006. 72 p. Acad. Diss. G 45. Roy, Himadri. Vascular Endothelial Growth (VEGFs) - Role in Perivascular Therapeutic Angiogenesis and Diabetic Macrovascular Disease. 2006. 81 p. Acad. Diss. G 46. Räty, Jani. Baculovirus surface modifications for enhanced gene delivery and biodistribution imaging. 2006. 86 p. Acad. Diss. G 47. Tyynelä, Kristiina. Gene therapy of malignant glioma. Experimental and clinical studies. 2006. 114 p. Acad. Diss. G 48. Malm, Tarja. Glial Cells in Alzheimer's Disease Models. 2006. 118 p. Acad. Diss. G 49. Tuunanen, Pasi. Sensory Processing by Functional MRI. Correlations with MEG and the Role of Oxygen Availability. 2006. 118 p. Acad. Diss. G 50. Liimatainen, Timo. Molecular magnetic resonance imaging of gene therapy-induced apoptosis and gene transfer: a role for 1H spectroscopic imaging and iron oxide labelled viral particles. 2007. 81 p. Acad. Diss. G 51. Keinänen, Riitta et al. (eds.). The first annual post-graduate symposium of the graduate school of molecular medicine: winter school 2007. 2007. 65 p. Abstracts. G 52. Vartiainen, Suvi. Caenorhabditis elegans as a model for human synucleopathies. 2007. 94 p. Acad. Diss. G 53. Määttä, Ann-Marie. Development of gene and virotherapy against non-small cell lung cancer. 2007. 75 p. Acad. Diss. G 54. Rautsi, Outi. Hurdles and Improvements in Therapeutic Gene Transfer for Cancer. 2007. 79 p. Acad. Diss. G 55. Pehkonen, Petri. Methods for mining data from genome wide high-throughput technologies. 2007. 91 p. Acad. Diss. G 56. Hyvönen, Mervi T. Regulation of spermidine/spermine N'-acetyltransferase and its involvement in cellular proliferation and development of acute pancreatitis. 2007. 79 p. Acad. Diss. G 57. Gurevicius, Kestutis. EEG and evoked potentials as indicators of interneuron pathology in mouse models of neurological diseases. 2007. 76 p. Acad. Diss. G 58. Leppänen, Pia. Mouse models of atherosclerosis, vascular endothelial growth factors and gene therapy. 2007. 91 p. Acad. Diss.


Recommended